Download as pdf or txt
Download as pdf or txt
You are on page 1of 344

1 | Introduction

ESTIMATING ROCK
MASS PROPERTIES

Robert Bertuzzi i
ESTIMATING ROCK MASS PROPERTIES

First Published 2019


Copyright © Robert Bertuzzi

This book is copyright. Apart from any fair dealing for the purposes of private study, research,
criticism or review, no part may be reproduced, stored in a retrieval system, or transmitted in any
form or by any means, electronic, mechanical, photocopying, recording or otherwise, without
the prior permission of the publisher

Published in Australia in 2019 by Robert Bertuzzi


PSM, Unit G3 56 Delhi Rd North Ryde New South Wales Australia

ISBN 978-0-6484420-1-1

ii
1 | Introduction

ESTIMATING ROCK
MASS PROPERTIES

Robert Bertuzzi

iii
ESTIMATING ROCK MASS PROPERTIES

Foreword

An undergound excavation designer works with This book is written from, and directed
limited resources: time and budget, and generally towards, a practitioner’s view. It is essentially
with very limited data. The designer needs robust the authors PhD thesis:
guidelines for selecting intact rock strength criteria, Bertuzzi R. Rock mass properties for tunnelling.
for assessing fracture stiffness, for classifying a rock PhD Thesis, School of Civil and Environmental
mass and for estimating rock mass parameters suitable Engineering, UNSW Australia, 2017.
for their design.
Throughout this book, the science of rock mechanics
is examined, critiqued, and distilled within the overall
aim of constructing guidelines with the specific aim
of improving the estimate of rock mass strength and
elastic parameters for underground design.

iv
1 | Introduction

Contents

Foreword iv 3.3.1 Tight defects 22


Contents v 3.3.2 Defects with thick infill 22
List of Tables viii 3.3.3 General defects 22
List of Figures x 3.3.4 Summary 24
Appendices xv 3.4 Hawkesbury Sandstone and
Symbols xvi Ashfield Shale 24
3.4.1 Shear strength 25
1 INTRODUCTION 1
3.4.2 Stiffness 26
1.1 Format 1
3.4.3 Sheared length 27
2 INTACT ROCK STRENGTH 3 3.5 Conclusions 27
2.1 Outline 3 3.5.1 Strength 27
2.2 Two dimensional criteria 5 3.5.2 Stiffness 29
2.2.1 Mohr-Coulomb 5
2.2.2 Griffith and Modified
4 ROCK MASS CLASSIFICATION 31
Griffith 5 4.1 Outline 31
2.2.3 Bieniawski 6 4.2 Q 31
2.2.4 Barton 6 4.3 GSI 35
2.2.5 Hoek & Brown 6 4.3.1 Chart 36
2.2.6 You 8 4.3.2 Quantification 39
4.4 Limitations 41
2.3 Three dimensional criteria 8
4.5 Conclusions 43
2.3.1 Drucker-Prager 8
2.3.2 Modified Wiebols & Cook 8 5 ROCK MASS STRENGTH CRITERIA 45
2.3.3 Lade 9 5.1 Outline 45
2.3.4 Modified Lade 11 5.2 Massive brittle rock 45
2.3.5 Extension strain 11 5.3 Coal 47
2.3.6 3D Hoek-Brown 11 5.3.1 A brittle rock 47
2.3.7 Christensen 12 5.3.2 Coal pillar strength 48
2.3.8 You 13 5.4 Plane of weakness 51
2.3.9 Other 13 5.5 Barton shear strength 52
2.4 Comparisons 14 5.6 Hoek & Brown shear strength 52
2.4.1 Colmenares & Zoback 14 5.6.1 Criterion 52
2.4.2 Kwasniewski 14 5.6.2 Limitations 54
2.4.3 Curve fitting 16 5.7 CWFS 56
2.5 Computer simulation 17 5.8 SRM 56
2.6 Conclusions 17 5.8.1 Calibration 57
5.8.2 Limitations 57
3 DEFECT PARAMETERS 19
5.9 Conclusions 58
3.1 Outline 19
3.2 Strength 19 6 ROCK MASS MODULUS 59
3.2.1 Patton 19 6.1 Outline 59
3.2.2 Barton-Bandis 20 6.2 Modulus Ratio 60
3.2.3 Summary 21 6.3 Data 62
3.3 Stiffness 22 6.3.1 Reliability 62

v
ESTIMATING ROCK MASS PROPERTIES

Contents

6.3.2 Sources 63 8.3 Results 113


6.4 Classification systems 63 8.3.1 Mostyn & Douglas 113
6.4.1 RQD 63 8.3.2 Artificial 119
6.4.2 RMR 64 8.3.3 Tunnelling Projects 119
6.4.3 Q 69 8.3.4 Colmenares & Zoback 119
6.4.4 GSI 69 8.3.5 Kwasniewski 120
6.4.5 Other 70 8.4 Discussion 120
6.4.6 Collated data 73 8.5 Conclusions 121
6.4.7 Conclusion 74
9 GSI CHARTS 123
6.5 Equivalent continua 75
9.1 Outline 123
6.6 Compliant inclusions 76
9.2 Sydney 129
6.6.1 Geophysics 76
9.2.1 Hawkesbury Sandstone 129
6.6.2 Rock mechanics 76
9.2.2 Ashfield Shale 129
6.6.3 Application 77
9.2.3 Sydney rock mass
6.7 Illustrative cases 79
classification 129
6.7.1 Ashfield Shale & Hawkesbury
9.2.4 Field data validation 131
Sandstone 79
9.2.5 Classification 135
6.7.2 Lac du Bonnet Granite 82
9.3 Reefton 137
6.7.3 Classification systems 84
9.3.1 Geology 137
6.7.4 Equivalent continua 86
9.3.2 Field data 137
6.7.5 Compliant inclusions 87
9.3.3 Classification 138
6.8 Conclusions 89
9.4 Macraes 139
7 BACK-ANALYSIS OF ROCK MASS 9.4.1 Geology 139
MODULUS 91 9.4.2 Field data 139
7.1 Outline 91 9.4.3 Classification 141
7.2 Theory 91 9.5 Quantification of GSI 141
7.2.1 Isotropy 91 9.6 Conclusions 145
7.2.2 Anisotropy 92
7.2.3 Methodology 93
10 TUNNELS IN HAWKESBURY
SANDSTONE CASE STUDIES 147
7.3 Case Studies 94
10.1 Outline 147
7.3.1 M2 Tunnel 95
10.2 Upper Canal Tunnels 149
7.3.2 Castle Hill Crossover Cavern 97
10.2.1 Geology 149
7.4 Conclusion 98
10.2.2 Cataract 149
8 INTACT ROCK TEST DATA 99 10.2.3 Devines 155
8.1 Outline 99 10.2.4 Numerical modelling 156
8.2 Data 100 10.3 Malabar 161
8.2.1 Mostyn & Douglas 100 10.3.1 Description 161
8.2.2 Artificial 102 10.3.2 Numerical modelling 162
8.2.3 Tunnelling Projects 102 10.4 Sydney LPG Storage Cavern 165
8.2.4 Colmenares & Zoback 110 10.4.1 Description 165
8.2.5 Kwasniewski 113 10.4.2 Numerical modelling 166
8.2.6 Summary 113 10.5 Northside Storage Tunnel 169

vi
1 | Introduction

Contents

10.5.1 Description 169 12.2.2 Limestone 250


10.5.2 Numerical modelling 169 12.2.3 Iron ore 254
10.6 Cross City Tunnel 174 12.2.4 Sandstone 256
10.6.1 Description 174 12.3 Laboratory samples 256
10.6.2 Numerical modelling 175 12.3.1 Stripa granite 256
10.7 Lane Cove Tunnel 177 12.3.2 Lac du Bonnet granite 259
10.7.1 Description 177 12.3.3 Artificially jointed granite 260
10.7.2 Numerical modelling 177 12.3.4 Quartzite 262
10.8 Results 180 12.3.5 Phyllite 262
10.9 Discussion 187 12.3.6 Other 265
10.10 Conclusions 187 12.4 Ok Tedi core testing 266
12.4.1 GSI 266
11 COAL PILLAR CASE STUDIES 191 12.4.2 Intact rock properties 267
11.1 Outline 191
12.5 Disturbance Factor 269
11.2 Data 191
12.6 GSI relationships 270
11.2.1 Australia 191
12.6.1 Validation 274
11.2.2 South Africa 194
12.7 Conclusions 277
11.2.3 USA 206
11.2.4 India 211 13 DISL – COAL 279
11.2.5 Combined 213 13.1 Outline 279
11.2.6 Other 214 13.2 Background 280
11.3 UCS 216 13.3 Conclusions 290
11.3.1 Australia 216
11.3.2 South Africa 216
14 REFERENCES 291
11.3.3 USA 217 APPENDICES 300
11.3.4 India 217
11.3.5 UK 217
11.3.6 China 217
11.3.7 Conclusion 217
11.4 Numerical modelling 219
11.4.1 Parameters 219
11.4.2 Generic cases 219
11.4.3 Australia 226
11.4.4 South Africa 229
11.4.5 USA 233
11.4.6 India 237
11.5 Discussion 238
11.6 Conclusions 243

12 ROCK MASS STRENGTH 245


12.1 Outline 245
12.1.1 Variability 245
12.2 Mine pillars 246
12.2.1 Metamorphic 246

vii
ESTIMATING ROCK MASS PROPERTIES

List of tables

1.1 Rock mass types 2 7.1 


Ea /E values for various E/E’ and G/G’
ratios at θ = 0 and 90° 94
2.1 Strength criteria for intact rock or
rock mass 4 8.1 The rock types that are contained
2.2 Values of A and B 10 within the laboratory test results
database 100
2.3 Poly-axial data used by various
researchers 15 8.2 The datasets with more than 20 test
2.4 Hoek-Brown parameters from results 101
best fitting curves 16 8.3 Artificial intact rock data set 102
8.4 The r2 and SEE [MPa] values obtained
3.1 Surface roughness contribution 20 from curve-fitting 118
3.2 Examples of joint stiffness values 24 8.5 The r2 values obtained from
curve-fitting the artificial test data 119
8.6 The r2 values obtained from
4.1 Case records of rock types used to
curve-fitting the data from
develop the Q system 32
tunnelling projects 119
4.2 The ESR values 34
8.7 The misfit values 120
4.3 Parameters used in RMR 35
8.8 The correlation obtained from
4.4 Components of JCond89 36 Kwasniewski (2013) 120
4.5 Terms to describe large-scale
waviness Jw 40
9.1 Sydney Classification System 129
4.6 Terms to describe small-scale
9.2 Suggested conditions for tunnelling
roughness JS 40
projects 130
4.7 Terms to describe joint alteration
9.3 Variability in strength and spacing as
factor JA 40
recorded in the drill hole database 133
9.4 Variability in the defect
5.1 Massive, brittle rock failure criterion 45 characteristics as recorded in the
5.2 Original ‘rock mass’ database 53 drill hole database 113
9.5 Typical and adverse conditions of
6.1 The tests carried out to assess the defects in Hawkesbury Sandstone 134
rock mass modulus 64 9.6 Typical and adverse conditions of
6.2 Case histories 65 defects in Ashfield Shale 134
6.3 In situ results from plate bearing tests 66 9.7 Reefton rock mass classes 138
6.4 Goodman borehole jack tests results 67 9.8 Macraes rock mass classes 140
6.5 In situ modulus collated 72 9.9 Comparison of quantified GSI and
6.6 Typical Parameters 83 that obtained from its chart 142
6.7 Rock mass modulus predicted from 9.10 Comparison of the GSI quantified
correlations UCS, RQD, RMR, Q following the method of Cai et al 145
and GSI 86
6.8 Prediction of rock mass modulus using 11.1 Summary of the UNSW Coal Pillar
Kulhawy and Zhang equations 87 Database 192
6.9 Prediction of rock mass modulus using 11.2 Details of the UNSW Coal Pillar
Hudson and David & Zimmerman 88 Database 193

viii
1 | Introduction

List of tables

11.3 Summary of the South African coal 12.4 Summary of RMR of NIOSH limestone
pillar database 194 mine database 251
11.4 Details of the South African Coal 12.5 Summary of iron ore pillar collapses 254
Pillar Database – Collapsed cases 195 12.6 RMR for the large granite core 257
11.5 Details of the South African Coal Pillar 12.7 Laboratory testing of 52 mm core
Database – Un-collapsed cases 198 of the Stripa granite 258
11.6 Summary of the USA coal pillar 12.8 Summary of other laboratory testing
database 206 of 52 mm core of Stripa granite 258
11.7 Details of the USA coal pillar 12.9 Summary of the UCS for the
database – Collapsed cases 207 intact rock and for the rock mass 265
11.8 Details of the USA coal pillar 12.10 Hoek-Brown parameters for the
database – Un-collapsed cases 208 limestone and hornfels test results 267
11.9 Summary of the Indian coal pillar 12.11 Broad categories of the Disturbance
database 211 Factor, D, for the rock mass
11.10 Details of the Indian coal pillar databases 269
database 212 12.12 Summary of the data used 270
11.11 Summary of the combined 12.13 Comparison of proposed
coal pillar database 214 relationships 271
11.12 Summary of coal UCS data 218 12.14 Hoek-Brown criterion parameters
11.13 Summary of compressive strength based on the GSI relationships 274
of 1m3 coal samples 218
11.14 Summary of coal Hoek-Brown 13.1 Details of the coal pillar cases
parameters that have been which do not conform to the
used by others 218 DISL envelope 288
11.15 Summary of average FoS
from Abaqus modelling of
pillar geometries 221
11.16 FoS results from UNSW method 222
11.17 Recommended widths of 2 m high,
square pillars 222
11.18 Pillar widths recommended 223
11.19 Comparison of FoS 225
11.20 Subset of the South African coal pillar
database used in modelling 229
11.21 Subset of the USA coal pillar
database used in modelling 233

12.1 Hard rock pillar database 246


12.2 Mining dimensions of NIOSH
limestone mine database 250
12.3  UCS testing of NIOSH limestone
mine database 250

ix
ESTIMATING ROCK MASS PROPERTIES

List of Figures

2.1  UCS test results of Sydney’s Hawkesbury 4.1 The original data set upon which
Sandstone 3 the Q system is based 31
2.2 Shear strength relationships for intact 4.2 The bolt spacing used in case histories 33
rock through to infilled defects 6 4.3 Q-chart 33
2.3 Comparison of direct tensile strength 4.4 Approximate relationship between
tests with BTS 7 Jr / Ja and apparent ‘shear strength’ 34
2.4 3D failure surface 9 4.5 Approximate relationship between Jr
2.5 Empirical relationship between m and η1 10 and JRC for 0.2 and 1 m long defects 35
2.6 Variation of the parameters η1 and m 4.6 Chart that linked a geological description,
with the ratio σc/σt 10 GSI and the Hoek-Brown criterion 36
2 7 Laboratory test results of 4.7 Original GSI chart 37
Blair Dolomite 13
4.8 Various rock mass problem scales 38
2.8 Conceptual strength envelope for
4.9 The changing sketches depicting the
intact rock 14
rock mass structure 38
2.9 Sari (2012) data with curve generated
4.10 Ranges of GSI for various qualities of
following Mostyn and Douglas (2000) 17
peridotite-serpentinite rock masses 39
2.10 Poly-axial test data 17
4.11 Quartzite rock mass downstream of the
Ord Diversion Dam, Western Australia 41
3.1 Theoretical bi-linear peak shear 4.12 Revised GSI chart 42
strength of a defect from Patton 19
4.13 Comparison of the range covered
3.2 Plot of JRCn versus JRC0 for joint by RQD 43
lengths varying from 1 m to 100 m 21
3.3 Relationship between peak dilation
angle 21 5.1 Correlation between estimated minimum
in situ strengths and crack initiation 46
3.4 Theoretical behaviour of a defect under
compression, tension and shear loading 22 5.2 Curved strength envelope 46
3.5 Influence factor, β, as a function of layer 5.3 Tensile crack initiation (CI) stress
thickness (h) and loading area (L, B) 23 versus UCS for various rock types 46
3.6 Experimental evidence for scale effect 5.4 Coal triaxial test data 47
on defect shear stiffness 23 5.5 Plan view of coal pillar geometry 49
3.7 Direct shear tests results for bedding 5.6 Measured pillar stress from various
and joints in Hawkesbury Sandstone 25 pillar geometries 50
3.8 Direct shear tests results for bedding 5.7 Variation of peak strength as a
and joints in Ashfield Shale 25 function of defect dip 52
3.9 Inferred shear stiffness for bedding and 5 8 Shear strength relationships 52
joints within Hawkesbury Sandstone 26 5.9 Relationships between RMR and the
3.10 Inferred shear stiffness for bedding parameters mb & s 53
and joints within Ashfield Shale 26 5.10 The influence of scale on the type of
3.11 Normal stiffness kn as function of rock mass behaviour model 54
bedding spacing 27 5.11 Variation of Hoek-Brown parameters
3.12 Shear stiffness versus defect length 28 for spall-prone rocks with mi = 20 55
3.13 The data from Figure 3.12 plotted 5.12 The range over which GSI can be used
over Figure 3.6 28 to predict the H-B strength parameters 55
3.14 The product ks•L 28

x
1 | Introduction

List of Figures

5.13 Cohesion-weakening and 6.23 Range of in situ modulus versus GSI 74


frictional-strengthening (CWFS) 56 6.24 Range of in situ modulus versus GSI
5.14 UDEC model for UCS tests on Lac du for disturbed rock masses 74
Bonnet granite and Aspo Diorite 57 6.25 Concept of equivalent elastic
5.15 Scaling of UCS values to block size 57 continuum 75
5.16  UCS tests on seven different 6.26 Theoretical behaviour of a defect 75
sedimentary rocks 57 6.27 Normalised effective bulk modulus of
6.1 Variation of intact Young’s Modulus a solid containing spheroids 77
to intact UCS 59 6.28 Correlation between RQD and Jv 78
6.2 Variation of modulus with shear strain 60 6.29 Comparison of the range covered
6.3 Intact modulus plotted against UCS 60 by RQD 78
6.4 Intact modulus plotted against UCS for 6.30 Typical photos of Ashfield Shale
Ashfield Shale, Hawkesbury Sandstone 61 Class II borehole core 80
6.5 Intact modulus plotted against UCS 6.31 Typical Ashfield Shale Class II exposed
data for Hawkesbury Sandstone 62 in the Castle Hill Crossover Cavern 80
6.6 The intact modulus plotted against 6.32 Typical photos of Hawkesbury
UCS data for Ashfield Shale . 62 Sandstone Class II borehole core 81
6.7 Correlation between RQD and Em/Er 63 6.33 Photo of saw cut Hawkesbury
6.8 Measured in situ rock mass modulus of Sandstone Class III becoming Class II 81
deformation against RMR76 64 6.34 Photo of the Lac du Bonnet granite 82
6.9 Measured in situ rock mass modulus of 6.35 Correlation between RQD and Em/Er 84
deformation against RMR76 65 6.36 Correlation between RMR and Em 84
6.10 Measured in situ rock mass modulus from 6.38 Correlation between GSI and Em 85
plate bearing tests against RMR76 66
6.39 Comparison of the various correlations 89
6.11 Deformation modulus obtained from
6.40 Comparison of the various correlations 90
Goodman borehole jacking tests 66
6.12 Deformation modulus obtained from
plate bearing tests 68 7.1 Circular hole in an infinite medium 91
6.13 Rock mass modulus and ratio of 7.2 Polar variation of Ea/E with the
rock mass to intact rock modulus 68 angle θ for various E/E’ ratios 93
6.14 In situ modulus against RMR for the 7.3 Example of Ea v K plot 93
Gotvand dam site 69 7.4 Example of an arbitrary excavation 93
6.15 Correlation between rock mass 7.5 Google map of Sydney 94
modulus and Q 69 7.6 Monitoring locations shown in the
6.16 Measured rock mass modulus of tunnel cross section 95
deformation against GSI 70 7.7 Results for Eh/Ev=1.0 96
6.17 Comparison of Equation 6.4 with 7.8 Section overview of cavern monitoring) 97
the rock mass modulus data 70
7.9 Results for G/G’=1.0 with E/E’=
6.18 In situ modulus versus GSI 71 0.7, 1.0, 1.5 and 2.0 180
6.19 Plate load test results of in situ
modulus against GSI 71 8.1 Data set from artificial laboratory
6.20 In situ modulus versus RMR76 71 testing programme 189
6.21 Comparison of measured and 8.2 Data set from the testing of
predicted in situ modulus 71 Hawkesbury Sandstone 190
6.22 Collated in situ modulus versus GSI 73 8.3 Data set from the testing of

xi
ESTIMATING ROCK MASS PROPERTIES

List of Figures

Brisbane Tuff 98 9.21 Stereoplots showing typical


8.4 Data set from the testing of Brisbane orientations at Macraes 140
phyllite 103 9.22 GSI values for Otago Schist rock mass
8.5 Data set from the testing of Brisbane at Macraes 141
quartzite 104 9.23 Variation in JCond for Hawkesbury
8.6 Data set from the testing of Brisbane Sandstone Class II 141
basalt 105 9.24 Comparison of quantified GSI and
8.7 Data set from the testing of Brisbane the GSI assessed from the chart 143
greywacke 106 9.25 Comparison of a quantified and
8.8 Data set for KTB amphibolite 107 chart GSI from Russo (2009) 143
8.9 Data set for Dunham dolomite and 9.26 Comparison of quantified and
Solenhofen limestone 108 chart GSI 144
8.10 Data set for Shirahama sandstone and 9.27 Distribution of logged spacing for
Yuubari shale 109 Hawkesbury Sandstone and
8.11 True 3D triaxial data sets 110 Ashfield Shale 144
8.12 Examples of the data used to
estimate the Lade parameters 111 10.1 Relationship of major horizontal stress
with depth and minor horizontal stress 148
10.2 Overall map of the Upper Canal 150
9.1 GSI chart for general rock mass and
schist 123 10.3 Lower hemisphere polar stereonet of
mapped joints in the Devines tunnels 151
9.2 GSI chart for flysch 124
10.4 Geological long section of
9.3 GSI chart for flysch 124
Cataract Tunnel 152
9.4 GSI chart for molasse at depth 125
10.5 Overbreak in the roof of the
9.5 GSI chart for molasse at surface 125 Cataract tunnel 153
9.6 GSI chart for ophiolites 126 10.6 Cataract tunnel 153
9.7 GSI chart for limestone 127 10.7 Cataract tunnel - Roof failure and
9.8 GSI chart for gneiss 127 fallout 154
9.9 GSI chart for the five rock masses 128 10.8 Inferred geological long section of
9.10 The wrong and right way to classify 130 Devines Tunnel 155
9.11 Equal area projections for Hawkesbury 10.9 Devines Tunnel No. 1 and No. 2
Sandstone and Ashfield Shale 131 Southern Portal 156
9.12 Subset of the data in Figure 9.11 131 10.10 Phase mesh for the Cataract tunnel 157
2

9.13 Distribution of logged spacing and 10.11 Phase2 mesh for the Devines tunnels 158
aperture for Hawkesbury Sandstone 132 10.12 Principal stresses predicted around
9.14 Distribution of logged spacing and the Cataract tunnel 159
aperture for Ashfield Shale 132 10.13 Principal stresses predicted around
9.15 Major horizontal stress versus depth 135 the Devines tunnel 159
9.16 GSI values for Hawkesbury Sandstone 10.14 Principal stress plot for the
rock mass 136 Cataract tunnel 160
9.17 GSI values for Ashfield Shale rock mass 136 10.15 Location of case studies around
9.18 Map of Reefton Goldfield 137 Sydney’s CBD 161
9.19 GSI values for Reefton greywacke 10.16 Malabar outfall tunnel 162
and argillite rock mass 138 10.17 Phase mesh for the Malabar ocean
2

9.20 Map of Macraes 139 outfall tunnel 163

xii
1 | Introduction

List of Figures

10.18 P
 rincipal stresses of the Malabar ocean 10.44 Principal and deviator stress versus 189
outfall tunnel 164 confining stress
10.19 Isometric view of the LPG 11.1 The database of Australian coal pillars 192
storage cavern 165 11.2 The database of South African
10.20 Failure in Gallery A 166 coal pillars 194
10.21 Phase mesh for the LPG storage
2
11.3 The database of USA development
caverns 167 workings coal pillars 206
10.22 Principal stresses of the first roof 11.4 The database of Indian coal pillars 211
collapse 168 11.5 The combined collapsed and
10.23 Alignment of the NST 169 un-collapsed database 213
10.24 Phase2 mesh for the Tunk’s Park to 11.6 Measured pillar stress from various
North Head tunnel 170 pillar geometries 214
10.25 Principal stresses of the 6.5 m 11.7 The databases superimposed onto
diameter Tunks Park to the plot in Figure 11.6 215
North Head drive 171 11.8 Close-up of Figure 11.7 215
10.26 Principal stresses of the drive 11.9 3D finite element model 220
under 95 m overburden 172
11.10 Contours of major principal stress
10.27 Principal stresses of the 3.8 m diameter and FoS 220
Tunks Park to Lane Cove drive 173
11.11 Definition of mining geometry 221
10.28 Alignment of the CCT 174
11.12 Comparison of the FoS result 223
10.29 Schematic long section 174
11.13 Finite element model and
10.30 Mapping of geology at the failure 175 material properties 224
10.31 Phase mesh for the CCT ventilation
2
11.14 Major principal stress for the
tunnel 175 case w =18.7 m and depth = 250 m 224
10.32 Principal stresses of the ventilation 11.15 Pillar load and inferred strength (Cp) 225
tunnel 176
11.16 Results of finite element modelling 225
10.33 Alignment of the LCT 177
11.17 Contours of pillar load and
10.34 Principal stress in the plane of the confining stress 227
modelled 184
11.18 Predicted σ1 v σ3 for the Australian
10.35 Phase mesh for the LCT tunnels
2
188 database 228
10.36 Principal stresses of the LCT 11.19 Example of the 2D FEM and
three-lane tunnel 189 material properties 230
10.37 Principal stress plot from the 11.20 Major principal stress for the
‘stable’ zones 181 case w =15.9 m and depth = 193 m 231
10.38 This is a subset of Figure 10.37 182 11.21 Predicted σ1-σ3 for the South African data-
10.39 Principal stress plot from the base 231
‘failed’ zones 183 11.22 Predicted σ1-σ3 of the South African
10.40 Deviator stress (σ1-σ3) versus confining collapsed pillar cases 232
stress (σ3) 184 11.23 Predicted σ1-σ3 of the South African
10.41 This is a subset of Figure 10.40 185 un-collapsed pillar cases 232
10.42 Deviator stress (σ1-σ3) versus confining 11.24 Major principal stress for the
stress (σ3) 186 case w =9.1 m and depth = 396 m 235
10.43 Difference between the criterion 11.25 Predicted σ1-σ3 of the USA collapsed
and the modelled principal stress 188 pillar cases 236

xiii
ESTIMATING ROCK MASS PROPERTIES

List of Figures

11.26 Predicted σ1-σ3 of the USA un-collapsed 12.19 Principal stress plots for the phyllite 264
pillar cases 236 12.20 Example of the core description,
11.27 Predicted σ1-σ3 plotted of the Indian photographs and sketches 266
pillar cases 237 12.21 Summary of UCS test results for
11.28 Predicted normalised principal stresses 238 hornfels and limestone 267
11.29 Predicted normalised principal stresses 239 12.22 Summary of UCS and TXL test results
11.30 Predicted normalised principal stresses for hornfels and limestone 268
for the collapsed pillars 240 12.23 Guidelines for estimating disturbance
11.31 Predicted normalised principal stresses factor, D 269
for the un-collapsed pillars 241 12.24  Proposed relationship between
11.32 Same data as shown in Figure 11.29 mb/mi and GSI 272
but as deviatoric stress 242 12.25 Proposed relationship between
11.33 The combined database of 162 s and GSI 272
collapsed pillars 243 12.26 Proposed relationship between
a and GSI 273
12.1 Ratio of σ1-σc versus w /h for the 12.27 Proposed relationship between
hard rock pillar database 247 mb/mi v GSI and s v GSI. 273
12.2 Ratio of σ1-σc versus average pillar 12.28  Normalised principal stress plots
confinement 247 for hornfels 275
12.3 Relationship between Cpav and w/h 248 12.29 Normalised principal stress plots
for limestone 276
12.4 Non-dimensional principal stress
σ1-σc versus σ3-σc 248
12.5 Non-dimensional principal stress 13.1 Normalised principal stresses 279
σ1-σc versus σ3-σc 249 13.2 Idealised strength envelope 280
12.6 Distribution of RMR in the 13.3 Principal stress plots of the
NIOSH database 251 collapsed coal pillar database 282
12.7 Summary of pillar layouts 252 13.4 Principal stress plots of the
12.8 Ratio of σ1-σc versus σ3-σc for the un-collapsed coal pillar database 283
limestone pillars 253 13.5 Principal stress plots of the
12.9 Ratio of σ1-σc versus σ3-σc for the coal pillar database 284
failed iron ore pillar database 254 13.6 Principal stress plots of the
12.10 Photos of Lorraine iron ore mines 255 collapsed coal pillar database 285
12.11 Ratio of σ1-σc versus σ3-σc for the 13.7 Principal stress plots of the
Laisvall mine test pillars 256 un-collapsed coal pillar database 285
12.12 Fracture traces mapped in the 13.8 Principal deviatoric stress plots of
large diameter Stripa granite core 257 the collapsed coal pillar database 286
12.13 Principal stress plot for the Stripa 13.9 Principal deviatoric stress plots of the
granite core 258 un-collapsed coal pillar database 286
12.14 Principal stress plot for the 13.10 Principal stress plots of the
Lac du Bonnet granite 259 collapsed coal pillar database 289
12.15 Samples of artificially jointed granite 260 13.11 Principal stress plots of the
un-collapsed coal pillar database 289
12.16 Results of laboratory testing of
artificially jointed granite 261 13.12 Recommended extended DISL
envelope for coal pillars 290
12.17 Principal stress plots for the quartzite 263
12.18 Digitised points of the phyllite 263

xiv
1 | Introduction

Appendices

ROCK MASS PROPERTIES FOR TUNNELING

CONTENTS
A Development of Barton shear strength criterion 301

B Development of Hoek-Brown strength criterion 304

C Development of Lade criterion 307

D Development of Christensen criterion 320

xv
ESTIMATING ROCK MASS PROPERTIES

Symbols

The symbols used in this book are listed below. While the intention is to respect and follow the original
source for the definition of symbols, where this would lead to confusion, alternatives are used. Multiple
definitions for the one symbol are separated by a semi-colon.
Throughout compression is assumed to be positive.

a r adius; Hoek-Brown criterion JA , JC , JS , JW joint alteration, joint condition


parameter; Lade criterion factor (= JW x JS /JA), joint small-
parameter scale roughness and joint large-
scale waviness, respectively
A area; variable
JV volumetric joint count
b, B width; variable
JCond joint condition factor of RMR
BTS Brazilian tensile strength
JRC, JCS joint roughness coefficient and
c Mohr-Coulomb criterion cohesion
joint wall compressive strength of
C variable Barton criterion
Co  iebols-Cook criterion
W k variable
parameter, UCS
kn , k s joint normal and shear stiffness,
Cp coal pillar strength respectively
Cpav average pillar confinement K  ulk modulus; representative
b
strength of 1 ft3 or m3 coal; stress
Cr subgrade reaction ratio
D disturbance factor K0 Yu criterion parameter
E Young’s modulus of elasticity L block size or joint length
Ei intact rock modulus m, mi, mb Lade criterion parameter; Hoek-
Ej modulus of the defect infill Brown criterion parameters
material M rock mass deformation modulus
Em rock mass modulus MR modulus ratio
f, f ’, f ’’ functions n  umber-density of cracks =
n
FoS Factor of Safety number of cracks / volume;
variable
G shear modulus
p vertical stress
GSI geological strength index
pa atmospheric pressure
h height of coal pillar; thickness
p0 pore pressure
H depth
q tan (45°+φ/2)
i dilation angle
Q rock mass quality (= RQD/Jn x Jr /Ja
I1, I3 stress invariants x Jw /SRF)
I1, I2 , I3 , I4 shape or influence functions Q’ t runcated rock mass quality (=
RQD/Jn x Jr /Ja)
Is(50) point load index
Q 0 , Q∞, QD Yu criterion parameters
J2 second deviatoric stress invariant
r radial distance
Ja, Jn, Jr, Jw joint alteration, number of joint
sets, joint roughness and joint r2 correlation coefficient
water factor, respectively

xvi
1 | Introduction

Symbols

R r atio of effective pillar width to γ unit weight of rock; Yu criterion


pillar height (= w/h or w1sinθ/h) parameter
RMi rock mass index Γ crack density
RMR rock mass rating: RMR’76 is the η1 Lade criterion parameter
1976 version; RMR’89 is the 1989
κ ‘mixed-mode’ fracture coefficient;
version
Christensen criterion parameter
RQD rock quality designation
friction coefficient of the micro-
μ
s average spacing of joints; Hoek- crack
Brown criterion parameter
ν Poisson’s ratio
su undrained shear strength
pi = 3.14159
π
Sp pillar load
Lode angle; variable angle
θ
SRF Stress Reduction Factor
Θ 2w2 / (w1+w2)
T0 unconfined tensile strength
σ1, σ2 , σ3 principal stresses (major,
TXL triaxial compressive strength intermediate and minor,
respectively)
ur, u Θ radial and tangential
displacement, respectively σc unconfined compressive strength
UCS uniaxial or unconfined σcrack normal stress required to close a
compressive strength crack
UCS* unconfined compressive strength σm mean principal stress
at which cracking initiates
σm,2 mean normal stress = average of
UTS uniaxial or unconfined tensile major and minor principal stress
strength
σn normal stress
Vb block volume
octahedral normal stress
σoct
Vp seismic compressive wave velocity
conventional triaxial strength
σs
Vs seismic shear wave velocity
unconfined tensile strength
σt
w width of coal pillar
σv, σH, σh vertical, major and minor
w1, w2 width and length of coal pillar, w1 horizontal in situ stress,
≤ w2 respectively
α aspect ratio of crack; Christensen σx, σy, σz triaxial stresses
criterion parameter
τ shear stress
β influence factor; angle which
τoct octahedral shear stress
the failure plane makes with the
major principal stress; (1–ν⁄2)κ ξ variable
δ displacement ψ equivalent friction angle
ε3 extension strain ´ superscript denotes effective
stress conditions
φ Mohr-Coulomb criterion angle of
friction i subscript denotes intact rock
parameter
φb, φc base friction angle
φr residual friction angle d subscript denotes defect
parameter

xvii
ESTIMATING ROCK MASS PROPERTIES

xviii
1 | Introduction

1 Introduction

Estimating strength and deformation parameters, impossible to test rock masses at a scale large enough
in particular the elastic modulus, of rock masses is and commensurate to most engineering applications.
fundamental to rock mechanics. It is at the core of Hence, estimates of the parameters are based mainly
analysis used for the design of slopes, foundations on empiricism.
and underground excavations in rock (Brown 2008). However, the database from which the empirical
However, the estimate of rock mass strength and methods are derived and against which the solutions
modulus remains the most debated issue between are tested is rather limited in terms of large rock
practitioners of rock mechanics. Because of a rock masses. There is a clear need to expand this database
mass’ very nature, inherent in any set of parameters are and scope to refine the approach to estimate the
various assumptions regarding: strengths and deformation parameters of rock masses
• rock type (lithology) for underground excavation design.
• strength of the rock substance (intact rock strength) The approach taken is to:
• fracturing of the rock mass by defects (bedding, • First, collate strength and deformation data relating
joints, shears, etc.) to large scale rock masses.
„„ persistence and spacing of defects • Second, to use this database to appraise the
„„ number of defect sets performance of existing criteria and relationships
„„ infill material in predicting the performance of rock masses in
„„ roughness of defects underground excavations.
• groundwater pressures • Third, to recommend robust methods to estimate
• reactivity of the rock to environmental change rock mass strength and modulus for the designer.
(shrink / swell and slake).
In addition, the strength and modulus of a rock 1.1 FORMAT
mass also depend on the type of problem being Jaeger in Paterson (1982) recounted the important
assessed and its scale. As the following examples show, things to study in rock mechanics are “(i) the properties
paraphrasing Stacey (2000), it is important to “get the of joints and (ii) the fundamental building blocks [are]
failure mechanism right”. the irregular pieces of rock between the joints which
• If the problem is a block driven by gravity, then would be very irregularly loaded”. This book is therefore
perhaps its strength is best described as defect shear arranged into two parts.
strength following the Mohr-Coulomb criterion.
The first is a critique of the literature in the areas
• If the problem is the fracture of either an intact that lead to estimates of rock mass properties, namely
sample of rock or a massive brittle rock mass, then intact rock and rock defects. The critique is extensive
perhaps the unconfined compressive strength (UCS as it is important to understand the science behind
or σc) or critical extension strain is a better measure. the problem before offering recommendations to the
• f the problem is failure of a large fractured, jointed
I
 practitioner.
rock mass body then a shear strength measure that
An extensive database of intact rock and large
encompasses the geological structure of the rock
scale rock mass failure is collated in the second part.
mass is appropriate.
The database is then used to test, compare, critique
Table 1.1 shows the range of rock mass problems and suggest alternative methods to estimate rock
that were considered by Hoek (1994) (1). mass properties.
To a great extent, the debate on rock mass
parameters stems from the fact that it is practically

(1)
It is important to note that Table 1.1 ignores sheared rock masses; an oversight corrected in later publications as will be discussed in subsequent chapters.

1
ESTIMATING ROCK MASS PROPERTIES

TABLE 1.1 Rock mass types; from by Hoek (1994)

Description Strength Strength testing Theoretical


Characteristics considerations

Intact rock Brittle, elastic and Triaxial testing of Behaviour of


generally isotropic core specimenas elastic isotropic
behaviour relatively simple rock is adequately
and inexpensive understood for
and results are most practical
usually reliable applications

Intact rock with Highly anisotropic, Triaxial tests Behaviour of


a single inclined depending on difficult and defects adequately
defect shear strength and expensive. Direct understood for
defect inclination shear tests most practical
preferred. Careful applications
interpretation of
results required

Massive rock with Anisotropic, Laboratory testing Behaviour of


a few defect sets depending very difficult complex block
on number, because of sample interaction in
orientation and disturbance and sparsley jointed
shear strength of equipment size rock masses poorly
defects limitations understood

Heavily jointed Reasonbaly Triaxial testing Behaviour of


rock masses isotropic, highly of representative interlocking
dilatant at low samples extremely angular pieces
stress levels with difficult because poorly understood
particle breakage of sample
at high stress disturbance
levels

Compacted rock Reasonably Triaxial testing Behaviour


fill or weakly isotropic, less simple but reasonably well
cemented dilatant and lower expensive due to understood from
congolmerates strength than large equipment soil mechanics
insitu rock due required to studies on
to destruction of accommodate granular materials
fabric samples

Loose waste rock Poor compaction Triaxial or direct Behaviour of


or gravel and grading allow shear testing loosely compacted
particle movement simple but waste rock and
resulting in expensive due gravel adequately
mobility and low to large size of understood for
strength equipment most applications

2
1 | Introduction

2 Intact Rock
Strength

UCS (MPa)

2.1 OUTLINE 0
0 10 20 30 40 50 60 70 80

Much of what we know about intact rock strength has


come from three relatively straightforward laboratory 10

index tests:
• the uniaxial tensile strength (UTS) test, which 20

for practical purposes is almost always replaced


by the Brazilian tensile strength (BTS) test (Pells 30

1993), though the BTS overestimates the UTS by Depth (m)

10 to 30% (Perras and Diederichs 2014, Read and 40


Richards, 2015)
• the uniaxial or unconfined compressive strength 50
(UCS) test
• the triaxial compressive strength (TXL) test. 60

Of these three, it is the UCS, which Jaeger (1966)


described as ‘the oldest, simplest and still the most 70
important test’, that remains the basic laboratory test
for rock mechanics studies. True 3D triaxial, or poly- 80
axial, tests are typically limited to specialised research.
Much has been written of the various testing methods, FIGURE 2.1 UCS test results of Sydney’s Hawkesbury
the tests’ reliabilities, the source of errors in the test Sandstone show an example of the variability usually seen
results, correlations between the tests’ results, etc. in laboratory testing programmes. The data has come from
investigations from three underground rail projects – Epping
Pells (1993) provides a good summary of the testing
to Chatswood Rail Link, CBD Metro and the Northwest Rail
and gives practical guidelines in carrying out UCS and Link – and a road tunnel – Lane Cove Tunnel.
BTS tests. He notes that errors in the testing come
from two sources “(i) bias in sample selection and (ii) expressions (or criteria) to model the strength of rock
errors resulting from inappropriate sample preparation, under different stresses.
test apparatus or test procedure”. The variability in
test results, which is seen in all testing programmes No attempt is made to cover the numerous criteria
such as that shown in Figure 2.1, stem not only that have been proposed for intact rock strength. The
from these errors but mainly from the inherent interested reader is directed to the works of Edelbro
variability of the intact rock itself. Further, the tests (2003) and You (2011) for example, for appraisals
themselves are not necessarily material properties but of the many criteria presented in Table 2.1. Of the
vary due to specimen geometry, grain size, inherent criteria listed, those most commonly adopted in rock
microscopic fractures, grain size etc. (Fairhurst 1971; mechanics practice and several of the more recently
Hardy et al. 1973). proposed are described in this chapter. These criteria
include Mohr-Coulomb, Griffith, Drucker-Prager,
Superimposed onto this data variability is the fact Wiebols-Cook, Hoek-Brown, extension strain,
that there is unlikely to be a general failure mechanism Bieniawski, Lade, Christensen and You. A brief
which is true for all rock types under all stress regimes. introduction to computer simulation of intact rock
For example, consider the different behaviour under testing is also presented.
tensile and compressive stresses (Fairhurst 1971) and
the very different visco-plastic behaviour of rock salt to The criteria are written as functions of either shear
the brittle behaviour of very high strength basalt. Yet and normal stress, τ=f(σn); or principal stresses, σ1=f(σ2,
in rock mechanics in the general and in underground σ3). Throughout this book compression is assumed to
design in the specific, there is a need for mathematical be positive as is the convention in rock mechanics.

3
ESTIMATING ROCK MASS PROPERTIES

TABLE 2.1 Strength criteria for intact rock or rock mass, after Edelbro (2003)

Strength Criterion Author

Coulomb (1776)
Mohr (1882)

Griffith (1924)

Balmer (1952)
Sheorey et al (1989)

Drucker-Prager (1952)

Hobbs (1964)

Fairhurst (1964)

Murrell (1965)

Franklin (1971)

Barton (1976)

Hoek & Brown (1980)

Stacey (1981)

Bieniawski (1974)
Yudhbir et al (1983)

Lade (1977)
Kim & Lade (1984)

Yoshida (1990)

Wiebols & Cook (1968)


Zhou (1994)

Zhang & Zhu (2007)

Christensen (2007)

You (2009)

4
2 | Intact Rock Strength

2.2 TWO DIMENSIONAL 2.2.2 Griffith and Modified Griffith


The premise of these criteria is that fracture of brittle
CRITERIA materials starts as tensile stress concentrations at the
tips of minute, thin cracks. Griffith used total energy
2.2.1 Mohr-Coulomb considerations to calculate the conditions for a crack
The Mohr-Coulomb criterion remains the simplest to grow under plane stress (σ1 ≠ 0, σ2 ≠ 0, σ3 = 0) and
and probably the most widely used in rock mechanics, derived the following relationship (Griffith 1921):
based on the number of commercial computer programs
that implement it and the number of publications
that discuss it e.g. Al-Ajmi and Zimmerman (2005);
Mostyn and Douglas (2000). It assumes that failure
occurs on a plane due to the shear stress (τ) on that which can also be written as:
plane which at failure is a function of that plane’s
cohesion (c) and friction angle (φ) and the normal
stress that acts on that plane (σn), i.e.:

Here, the criterion is written with compression as


positive and T0 = - σt. A detailed explanation of the
where c and φ are related to the unconfined compressive
criterion can be found in Jaeger and Cook (1979). The
and tensile strengths σc, σt as:
criterion ignores frictional forces between the cracks’
faces and always predicts σc = 8σt.
The Griffith crack theory was modified by several
researchers, in particular McClintock and Walsh
The criterion can also be expressed in terms of principal (1962) who included the effect of friction along
stresses as: the crack faces and Murrell (1963) who extended
the criterion into true triaxial stress conditions
(σ1 ≥ σ2≥ σ3).
McClintock and Walsh (1962) included the
effect of friction along the crack faces to derive the
following equation.

The Mohr-Coulomb criterion is particularly suited


to analysing the shear strength of planar features such
as joints and other defects. However, it is limited in its
ability to analyse the peak rock strength as: When the crack closes, this becomes equivalent to
the Mohr-Coulomb criterion. That is, provided the
• actual peak strength envelopes are generally
normal stress required to close a crack (σcrack) is small
non-linear
enough to be neglected then the equation becomes:
• it implies failure occurs as shearing and in the
direction, π/4 + φ/2
• it implies c and φ are mobilised simultaneously,
i.e. it is fundamentally flawed for brittle materials
(Barton and Pandey 2011; Christensen 2007)
• it implies the intermediate principal stress (σ2) has This is the same as the Mohr-Coulomb criterion but
no effect on failure with σt = 0.5c.
• it predicts the ratio of unconfined compressive to Murrell (1963) extended the criterion into true
tensile strength, σc/σt, is always equal to (1+sinφ)/ triaxial stress conditions with the following Modified
(1-sinφ), which is much less than experimental Griffith equation:
evidence. For example, for the range of φ = 20
to 50°, σc = (2 to 7.5)σt, whereas typical ratios are
between 10 to 20 (Lade 1993). To counter this,
a tension cut-off is usually included as part of the
criterion for rock mass strength.

5
ESTIMATING ROCK MASS PROPERTIES

While some research provides compelling evidence The general form of the Barton shear strength
to show that general shear failure in rock may actually criterion, Equation 2.2, has been proposed for the
occur as tensile failure at the tips of developing cracks, strength of intact rock, jointed rock, rock fill and
e.g. Diederichs et al. (2004); Tarasov and Potvin infilled defects as shown in Figure 2.2, e.g. Barton
(2013), the Griffith and Modified Griffith crack (1999); (2014).
extension criteria oversimplify the process of rock
failure. The inherent assumption in their formulation is
that the extension of a single crack defines rock failure.
The criteria were found to give poor agreement with
EQUATION 2.2
experimental results from multi-axial compression tests
on rock samples which led subsequent researchers, e.g.
Hoek and Brown (1980); to develop alternative forms. where X represents roughness; Y, the effective
compressive strength and Z, the basic and residual
Hoek and Martin (2014) go so far as stating that friction. For intact rock, X = 50, Y = and Z = φc
while the Griffith crack theory “ is very important = 26.6°. Consideration of this criterion is presented in
background material for an understanding of the Appendix A.
mechanics of brittle failure initiation, it is of limited
practical value in the field of rock engineering.”

2.2.3 Bieniawski
Stacey and Page (1986) quote the Bieniawski (1974)
relationship, which is written below, stating that
adopting A = 3.5 and k = 0.75 should not result in
errors between predicted and measured intact strength
of much greater than 10%. However, they note that
the criterion is limited to shear failure in massive,
non-brittle rock. Further, the fact that it is undefined
in tension (σ3 < 0) may restrict its usefulness as a
general criterion.

EQUATION 2.1
FIGURE 2.2 Shear strength relationships for intact rock
through to infilled defects (Barton 1999; 2014)
2.2.4 Barton
In the late 1960s and early 1970s, Barton researched
the strength and stiffness of rock joints e.g. Barton 2.2.5 Hoek & Brown
(1973). A fuller discussion of that work is provided in Hoek and Brown started with the Griffith crack
Chapter 3 which deals with defect shear strength and extension theory to derive their empirical failure
Chapter 5 which deals with rock mass strength. In this criterion, which is written below, by trial and error
chapter, the focus is on Barton’s fitting of an empirical (Hoek and Brown 1980). Their aim was to provide a
relationship to the published triaxial strength data simple, empirical rock failure criterion that adequately
of three intact rocks – Solenhofen limestone, Spruce describes the response of an intact rock sample from
Pine dunite and Westerly granite. He did this by uniaxial tensile stress to triaxial compressive stress.
adopting the Bieniawski criterion (Equation 2.1) to
fit the test data and developed an empirical method
to derive the shear (τ) and normal (σn) stresses from
it. The following relationship to estimate the shear
strength of the mobilised failure plane was proposed
(Barton 1976). Because of the criterion’s importance to rock
mechanics, it’s derivation from the Griffith crack
extension theory (Griffith 1921) is presented in
Appendix B and a full description of its development
is given below.
The criterion was based on an extensive intact
rock data set, which had a range of σc from 40 to

6
2 | Intact Rock Strength

580 MPa (Brown 2008)(2). Their triaxial data meant Figure 2.3). Perras and Diederichs (2014) found that
the intermediate principal stress, σ2, equalled the the σt / BTS ratio varies from 0.7 for sedimentary, 0.8
minor principal stress, σ3, and hence the criterion is for igneous and 0.9 for metamorphic rocks.
independent of σ2 . The criterion was subsequently
extended to rock mass failure based on a modest
collection of triaxial data of 6 inch diameter by 12
inch long core and re-compacted samples of Panguna
andesite (Hoek and Brown 1980). This is discussed in
Chapter 5.
For intact rock, the constants m and s take their
maximum values of mi and 1.0, respectively. Hoek
and Brown (1997) make the point that to be consistent
with the original formulation, reliable values for σc and
mi need to be estimated from laboratory triaxial tests
over the range 0 < σ3 < 0.5σc.
There are contradictory statements about the
empirical constants, m and s.
In their original work, Hoek and Brown (1980)
state that mi can be approximated by σc/|σt|. However,
Hoek (1983) stated ‘there is no fundamental relationship
between the empirical constants included in the criterion FIGURE 2.3 Comparison of direct tensile strength tests
[m and s] and any physical characteristics of the rock’. The with BTS, from Read and Richards (2015)
view that the Hoek-Brown parameter mi is not rock
type dependent is furthered in research subsequent Several limitations of the Hoek-Brown failure
to the equation’s derivation, e.g. Marinos and Hoek criterion were identified during its development
(2000); Mostyn and Douglas (2000). However, tables including that it is independent of σ2 and it was based
of mi relating to rock type are ubiquitous. They are only from test data that was inferred to be related
given in the original papers, Hoek and Brown (1980); to brittle failure as defined by σ1 > 3.4σ3 (Hoek and
(1980), appear in subsequent publications and in the Brown 1980). Note that this definition of the brittle-
widely used rock mechanics computer programs by ductile transition boundary used by Hoek & Brown
RocScience Inc (Canada). differs from that originally developed by Mogi (1966)
Further adding to the debate, Zuo et al. (2008); of (σ1- σ3) = 3.4σ3 and confirmed in Mogi (1972).
Zuo et al. (2015) derive a strength criterion in the same Later research identified further limitations with the
format as the original Hoek-Brown by considering Hoek-Brown failure criterion.
the development of micro-cracks in intact rock. In so
doing, they state that the parameter m, given here as
• 
Mostyn and Douglas (2000) highlighted the
following concerns.
equal to , can characterise the rock.
„„ The criterion is undefined for σ3 < - σc /mi, i.e. it
doesn’t predict the tensile strength of intact rock.
Hoek and Martin (2014) summarise research
confirming that the Hoek–Brown criterion does
not predict the tensile strength of intact rock and
For the values typical of hard rock, the ratio (see
suggest adopting a tensile cut-off.
derivation in Appendix B) suggests that the Hoek-
Brown parameter, m, is: „„ etter curve fitting is achieved if the parameter
B

a is variable instead of being a constant value of
0.5. A variable a also captures brittle and ductile
behaviour better. Others have noted this as well
(Carter et al. 2007; Carter et al. 2008; Diederichs
Hoek and Martin (2014) offer what they term a et al. 2007; You 2011).
‘preliminary relationship’ between mi and σc/|σt|, „„ The method of fitting the criterion to test data
namely: σc/|σt|| = 8.62 + 0.7mi. However, Read and affects the estimates obtained of the constants.
Richards (2015) confirm Hoek and Brown (1980) They recommended a least squares procedure to
original’s opinion that mi can be approximated by curve fit the criterion minimising the following
σc/|σt| and further note on the basis of a comparison errors, i.e.
of the published results of 66 pairs of direct tensile (measured σ1 – predicted σ1) for σ1 > -3σ3
strength and BTS, that σt is approximately 0.9BTS (see (measured σ3 – predicted σ3) x mi for σ1 ≤ -3σ3
(2)
The weakest intact rock was a sandstone with UCS = 39.9 MPa

7
ESTIMATING ROCK MASS PROPERTIES

„„  ey too concluded that mi can be approximated


Th Brown criteria on the basis of the mean misfit. It is
by σc/|σt| worth noting that the You criterion quickly over-
• K
 aiser et al. (2000) showed using back-analyses that predicts the tensile strength of rock. That is, for cases
the criterion was not well suited to brittle failure. where σ3 < 0, the exponential term in the criterion
This is an interesting observation as Hoek and becomes negative. In fact, You (2012) recommends
Brown (1980) only included data that related to that a separate polynomial equation is derived for
brittle failure in developing the criterion. the tensile strength criterion and You (2015) adopts a
• Barton and Pandey (2011) state that the failure tensile cut-off.
mode involving rock breakage is highly unlikely
to be governed by either the Mohr–Coulomb or
Hoek–Brown style of strength criterion and offered
2.3 THREE DIMENSIONAL
the “c then tanφ” approach e.g. Hajiabdolmajid et CRITERIA
al. (2002).
The generalised Hoek-Brown criterion introduced 2.3.1 Drucker-Prager
by Hoek (1994) and then Hoek et al. (2002) made the There are several poly-axial or true three-dimensional
parameter a variable between 0.5 and 0.65 for rock (3D) shear strength criteria which have been proposed
mass failure. to explicitly consider the influence of the intermediate
principal stress, σ2 . Perhaps the first commonly used
3D criterion for intact rock was the Drucker-Prager
criterion (Drucker and Prager 1952) which extended
the earlier von Mises criterion. The criterion can be
Further discussion of the Hoek and Brown criterion written as:
and its applicability to rock mass failure is presented
in the subsequent Chapter 5 which deals with rock
mass strength.

2.2.6 You where


You (2009) developed an empirical strength criterion by
curve-fitting an exponential function to conventional
triaxial stress test data (σ1 ≥ σ2 = σ3). The criterion
is written in terms of three independent material
and the octahedral stresses are defined in terms of the
parameters; Q 0 , Q∞ and K0 .
principal stresses as:

where

σs is the conventional triaxial strength at confining


pressure σ3 However, it has been shown that this criterion does
not match experimental intact rock data very well e.g.
Q 0 is the UCS, σc
Colmenares and Zoback (2002); Yu (2006).
Q∞ is the maximum deviatoric strength. This is
the same parameter which Barton (1976) defines 2.3.2 Modified Wiebols & Cook
as the rock’s confined strength (σ1-σ3) Wiebols and Cook (1968) following the Griffith crack
K0 is the rate of increase of strength to confining extension theory (Griffith 1921) developed a criterion
stress at σ3 =0 considering strain energy within intact rock.
They noted that the Griffith crack theory implied
The arbitrary parameters γ and QD are found by
that compressive failure starts and continues from the
least squares curve-fitting to test data.
weakest crack. However, they observed detailed tests
which showed that compressive failure starts “first from
It can be readily seen that the criterion fits an the most dangerous crack but increasing stress is required
exponential curve between the points Q 0 and Q∞. For to increase the length of this crack, in the course of which
σ3 = 0, the equation devolves to σs = Q0 = σc. As σ3 failures start from other, less dangerous, cracks.” They
becomes large, σs approaches Q∞. continue: “the most dangerous crack must lie in a plane
You (2009) assessed that his criterion better matches parallel to the direction of the intermediate principal
test data than either the Mohr-Coulomb or the Hoek- stress, but this is not necessarily the case for other cracks

8
2 | Intact Rock Strength

which subsequently extend”. These other cracks which 1 It is a smooth, triangular, monotonically curved
subsequently extend can be in any direction. failure surface in the octahedral plane
They postulated an effective energy criterion for the 2 The opening angle of the failure surface is described
strength of rock. Wiebols and Cook (1968) presented by the material’s friction angle, as in the Mohr-
a numerical method to evaluate the effective shear Coulomb criterion
strain energy of intact rock subjected to 3D stresses. 3 The failure surface is curved in planes containing
Zhou (1994) presented the Wiebols-Cook criterion the hydrostatic axis
in the following form, which has subsequently been 4 It can have a tensile strength.
referred to as the Modified Wiebols & Cook criterion,
e.g. Colmenares and Zoback (2002).

where the variables A, B and C are given as:

FIGURE 2.4 3D failure surface viewed (a) from the side to


show curved envelope in triaxial plane and biaxial planes
and (b) from the end to show smoothly rounded, triangular
shape in octahedral plane (Lade 1993)

The Lade criterion is usually written as:

2.3.3 Lade EQUATION 2.3

Lade originally proposed a semi-theoretical, 3D failure


criterion to describe the results of laboratory strength where
testing of cohesion-less soil (Lade and Duncan 1975).
Lade assumed that the cohesion-less soil, Monterey
No.0 Sand, behaved isotropically and hence wrote a
failure criterion in terms of the stress invariants I1 =
σ1 ≥ σ2 = σ3 and I3 = σ1 ≥ σ2 = σ3. Lade empirically
developed the criterion with further laboratory testing • The value of I13 / I3 is 27 at the hydrostatic axis σ1
on other cohesion-less soils: Sacramento River Sand, = σ2 = σ3
Crushed Napa Basalt and Painted Rock Material • The term apa reflects the tensile strength of the
(Lade 1977). material and is slightly greater than the UTS. Lade
(1993) uses values of a that are 0.1 to 2% higher
than σt/pa.
• The apex angle of the failure surface increases with
where κ1 is a constant whose value depends on the the value of η1. The curvature of the failure surface
density of the sand. increases with the value of m.
A translation of the principal stresses along the The relationship between the parameter used to
hydrostatic axis was later introduced to include the describe the opening angle of the failure surface, η1 and
effective cohesion and tension that can be sustained by that used to describe the curvature of the failure surface,
concrete and intact rock Kim and Lade (1984); Lade m, has been investigated by Lade (2013) who reported
and Kim (1995). good correlation using Equation 2.4 as listed in Table
Lade’s observations of laboratory testing of intact 2.2. The resulting plots are shown in Figure 2.5.
soil, rock and concrete under uniaxial, biaxial, triaxial
and torsional stresses, led him to conclude that a three-
dimensional failure surface should have four main
characteristics as shown in Figure 2.4. EQUATION 2.4

9
ESTIMATING ROCK MASS PROPERTIES

TABLE 2.2 Values of A and B for Equation 2.4 from Lade (2013)
Rock Type A B r2 % that Predicted σt
apa > σt [MPa]

Igneous 0.249 55.1 0.9265 13 440

Metamorphic 0.257 22.7 0.8330 28 277

Sedimentary 0.210 2.15 0.9011 156 875

Organic (coal) 0.358 130.9 0.9765 6 29

All rocks (no coal) 0.229 9.62 0.8752 56 642

Lade’s equation is developed in Appendix C. That work 2.3.3.1 Typical ranges of parameters
shows that for the five values of A and B in Table 2.2: The following equations, which are developed in
• the inferred values of apa are approximately 6 to Appendix C, show Lade’s parameters are mathematical
156% greater than σt functions of the compressive strength and tensile
• the predicted tensile strengths are extremely high strength of the rock.
and artificial.
The surprising and unexpectedly high values
of apa and σt found by the above analysis questions
the usefulness of the proposal to estimate the Lade
Specifically,
parameters from a single laboratory test as suggested
by Lade (1993).

The variation of the parameters m and η1 with


respect to the ratio is shown in Figure 2.6. This
graph suggests that for intact rock, which typically has
values of between 8 and 15, m ≈ 0.1 – 1.25 and η1 ≈
500 to 50000, that is slightly tighter ranges than those
suggested by Lade (1993).

1.00E+06 2.00
UCS= 1MPa
UCS= 10MPa
UCS= 100MPa
1.00E+05 1.50
UCS= 250MPa
m

η1 1.00E+04 1.00 m

1.00E+03 0.50

1.00E+02 0.00
0 2 4 6 8 10 12 14 16 18 20
σc/σt

FIGURE 2.5 Empirical relationship between m and η1; after FIGURE 2.6 Variation of the parameters η1 and m with the
Lade (2013) ratio σc/σt. The typical range for the ratio σc/σt is between 8
and 15, which implies that η1 typically lies in the range 500
to 50000 and m typically lies in the range 0.10 to 1.25.

10
2 | Intact Rock Strength

2.3.4 Modified Lade these levels of strain correspond to a stress level in


It is worth mentioning the Modified Lade criterion as the region of σ1 = 0.3σc. Diederichs (2007) observes
developed by Ewy (1999) as it too has been discussed failures around underground excavations do not
in literature as a 3D failure criterion. necessarily match the predictions from the extension
strain criterion. The relationship σ1 = 0.3σc is in
Ewy (1999) modified the Lade criterion to solve his
keeping with later research in the onset of fracturing in
particular problem of assessing mud weight to control
brittle rock. This is discussed in Chapter 5 in dealing
the wall stability of wellbores. Ewy chose to:
with rock mass strength.
• set m to a constant value of 0, i.e. a straight-line
failure envelope
2.3.6 3D Hoek-Brown
• replace η with η1 and made it a function of friction Several extensions to the Hoek-Brown equation to
angle of the rock φ
explicitly consider 3D stresses have been proposed. One
• replace apa with S1, and made it a function of of the earliest was that proposed by Pan and Hudson
cohesion and friction angle of the rock c, φ (1988) who re-wrote the Hoek-Brown equation in
to effectively obtain a criterion that predicts linear terms of stress invariants and explicitly included the
shear strength increase with increasing I1. Ewy intermediate principal stress when they proposed the
obtained a closed-form solution for critical mud weight following form.
applicable to arbitrary wellbore orientations and in situ
stress states.
The Modified Lade criterion is:
EQUATION 2.5

where Where I1 is the first stress invariant and J2 is the


second deviatoric stress invariant:

Substituting and noting that the octahedral stresses


are:

p00

As the strength envelopes of intact rock are strongly


curved and not linear as predicted by the Modified Equation 2.5 can be re-written as:
Lade criterion, this criterion is not considered further.

2.3.5 Extension strain


Stacey (1981) provided an extension strain failure
criterion for brittle rocks stating that rock fracture
initiates when strains exceed some critical extension
value. The criterion is applicable to the initiation of
fractures and not of rock failure per se, so strictly it is
not a strength criterion.
For a material with ideal linear deformation
behaviour, the extension strain is:

and hence

Stacey (1981) and Stacey and Page (1986) provide


values of critical extension strain for a few rock types,
typically in the range 160 to 300 μm, and note that EQUATION 2.6

11
ESTIMATING ROCK MASS PROPERTIES

Zhang and Zhu (2007) noted that Equation 2.6


does not simplify to the Hoek-Brown criterion in
conventional triaxial tests, i.e. when σ2 = σ3. They
instead proposed that the mean normal stress be
restricted to consider only σ1 and σ3. By not including
σ2 into the calculation of the mean normal stress, It is apparent that the limitations of the Hoek-Brown
Zhang and Zhu (2007) followed the observation made criterion except for σ2 independence, will apply to the
by Mogi (1967) that the initial failure in tested samples Melkoumian et al and the Zhang & Zhu extensions.
occurred in the σ2 direction and hence the normal Further, the limitations of Drucker-Prager are likely to
stress acting on that failure path was a function of only apply to the Melkoumian et al extension.
σ1 and σ3. The octahedral stress σoct in Equation 2.6
therefore could effectively be replaced by the mean 2.3.7 Christensen
normal stress σm,2 = (σ1+σ3)/2, i.e. Working with ceramics and metals, Christensen
developed a theoretical, two parameter, 3D strength
criterion for homogeneous, isotropic materials
(Christensen 1997; Christensen 2004; Christensen
2005; Christensen 2006; Christensen 2006;
EQUATION 2.7
Christensen 2007). The criterion is effectively a
non-linear form of the Drucker-Prager criterion
Equation 2.7 differs from Equation 2.6 in the mean (Christensen 2006).
stress term. Christensen’s initial objective was to generalise
Adopting mb = mi, s=1 for intact rock and assuming the von Mises criterion to include the effect of mean
values for mi, Zhang and Zhu (2007) compared normal stress. He wanted to (i) explicitly integrate a
Equation 2.7 with the 187 poly-axial test results fracture mechanism into the criterion and (ii) derive
from five data sets collated by Colmenares and a criterion that determines whether a failure mode in
Zoback (2002) and concluded that it gave reasonable any particular state of stress is expected to be ductile
correlations. or brittle.
The failure surface generated by Equation 2.7 Christensen started with a relationship between
however is not smooth and convex in all the stress the elastic energy and the dilatational and deviatoric
states, apparently a common problem for strength stresses within a body. Further, he assumed that
criteria based on Mogi (1967) that has the mean homogeneous, isotropic materials do not fail under
stress only considering σ1 and σ2 . To overcome this hydrostatic compression but could fail under
limitation Zhang et al. (2013) suggest re-arranging the hydrostatic tension. The criterion is written as
criteria to be a function of the Lode angle, θ and then follows, noting that here compression is positive as in
limiting θ to ensure the failure surface remains smooth standard rock mechanics, i.e. opposite convention to
and convex. Similar approaches have been used to Christensen’s original criterion. The criterion applies
ensure smooth and convex failure envelopes of other to materials that have a |UTS| ≤ |UCS|.
3D criteria, e.g. Sloan and Booker (1986).
In another approach, Melkoumian et al. (2009)
extended the work of Priest (2005) in developing an
explicit solution for a 3D Hoek-Brown criterion by
combining its 2D version with the 3D Drucker-Prager
criterion. They too compared their equation with the
EQUATION 2.8
data collated by Colmenares and Zoback (2002) and
similarly concluded that it gave reasonable correlations.
The Melkoumian et al effective yield stress, σ’zf, can
be found using the following equations at any major
principal effective stress point, σ’x and σ’y. The restriction σ1 ≤ |σt| if |σt| < ½ σc is the explicit
fracture criterion, i.e. brittle failure.
Christensen defines two material parameters, κ and
a as functions of UCS (σc) and UTS (σt):
where

12
2 | Intact Rock Strength

The parameter κ changes the scale of the strength This is effectively a non-linear form of the Drucker-
envelope; and the parameter α is non-dimensional and Prager criterion discussed in Section 2.4.1 (Christensen
controls the shape of the envelope. The special case of 2006). Its attraction is that it takes into consideration
α = 0, is the von Mises yield behaviour of very ductile 3D stress (σ1≥σ2≥σ3) yet needs only two parameters; κ
metals, i.e. those with |σt| = σc. The limiting case and α or really σt and σc. The parameter σt is difficult
of α→∞ is the brittle limit with the tensile strength to measure (Hoek and Brown 1980) it is usually
σt being negligible compared with the compressive approximated with the BTS.
strength σc. Many rocks and certainly most rock A series of equations are developed in Appendix
masses would fall into this category. D from the Christensen criterion showing that while
The ductile-brittle delineation is given by the the form of the Christensen criterion is similar to the
following relationship: Hoek-Brown criterion, it does not simplify to the
Hoek-Brown criterion under triaxial conditions (σ2
= σ3). This is a similar issue as the Pan and Hudson
(1988) 3D Hoek-Brown criterion (Equation 2.6).

2.3.8 You
You (2009) proposed an extension to his 2D criteria to
where the mean normal stress or octahedral stress is: consider true 3D stress (σ1 ≥σ2≥σ3) with the following
five parameter criterion.

Christensen appreciated that with only two material


parameters, κ and α, the criterion would not match
all test data on the variety of natural and man-made Where σs is given in the 2D version of the criterion
materials with high accuracy but suggested that the in Section 2.3.7 and the arbitrary parameters γ and QD
criterion would allow reasonably accurate models. For are found by least squares curve-fitting to test data.
example, the criterion for Blair dolomite compared The parameter γ could be a constant equal to about 1.7
to laboratory test data and Mohr-Coulomb is shown (You 2009; You 2012).
in Figure 2.7. It can be seen that the Christensen It is clear that σ1 reaches the maximum value
criterion matches the general trend of the data but does (σs+QD) when:
not replicate it exactly.

You (2009) states that at least six sets of triaxial


compression, triaxial extension and biaxial test data
are required to determine the three parameters, Q0,
Q∞, K0 – which are required for conventional triaxial
stress conditions – plus the further two parameters,
QD and γ - which are required for true 3D stress states.
Interestingly, UCS data, which are often the main if
not only test carried out (see this chapter’s introductory
FIGURE 2.7 Laboratory test results of Blair Dolomite paragraphs), are not to be used to estimate QD and γ
(reproduced from Christensen, 2004, which shows (You 2009; You 2012).
compression as negative). C-M is the Mohr-Coulomb
criterion.
2.3.9 Other
The criterion can be written in terms of the Singh et al. (2011) modified the Mohr–Coulomb
octahedral normal and shear stresses as: criterion to take into account non-linear triaxial and
poly-axial strength of intact rocks. Adopting the term
used by Barton (1976), they considered that beyond
the critical confining pressure (σcrit), which is based
on a statistical approach of published triaxial data
and hence in principal stress space as: they estimated to be approximately the UCS, there is
no further increase in the deviatoric stress at failure.
This was extended to include anisotropic rock masses
by Singh et al. (2015); Singh et al. (2015) with the

13
ESTIMATING ROCK MASS PROPERTIES

following expression. Perhaps not surprisingly they also found that the
Mohr-Coulomb and Hoek-Brown criteria fitted the
data at least equally well for those rocks which behaved
largely independent of σ2 . These were the two high
strength rocks in their database, which had a UCS of
95 and 120 MPa. These observations are in keeping
with the conclusion reached by Mogi (1967) who
stated that the effect of σ2 is “much more pronounced
for brittle than for ductile materials”. Two interesting
where σcβ is the UCS of the rock in a given direction observations of Colmenares and Zoback’s work have
β; c β0 and φβ0 are the Mohr-Coulomb parameters been made by others.
obtained from triaxial tests at low confining pressures;
σcrit is estimated to be approximately 1.25UCS and • A l-Ajmi (2006) found that the variance between
the equations and the data is due more to the
experimental scatter than to any inherent difficulty
in curve fitting the data.
• You (2009) argues that the seven criteria evaluated
Other poly-axial strength criteria, which have also by Colmenares and Zoback (2002) only correlated
been statistically derived by curve-fitting the same 187 with the experimental data because of the strong
test results used by Colmenares and Zoback (2002) influence of σ1 not because they capture the influence
have been recently proposed for intact rock, e.g. Jiang of σ2 and σ3 particularly well. Not surprisingly, You
et al. (2011); Mortara (2010); Rafiai (2011). (2011) confirmed that three parameter criteria (of
which the Lade, the generalised Hoek-Brown and
Rafiai et al. (2013) show that an artificial-neural-
the You criterion are some examples) can better fit
network formulation can be trained with part of a
test data than two parameter criteria (which include
database of poly-axial laboratory tests on intact rocks
the Mohr-Coulomb, the original Hoek-Brown and
and then used to successfully predict the rest of the
by inference the Christensen criterion).
database. Obviously at this stage, this approach is
limited to samples that can be extensively tested in • You (2011) ranked the Hoek-Brown criterion along
a laboratory but it may lead to estimating rock mass with his exponential criterion above the other 14
strength in the future. criteria he evaluated as better able to fit laboratory
test data.
There are several issues with these recently proposed
criteria in addition to their increasing complexity.
2.4.2 Kwasniewski
They are still at research stage and are not yet suitable
for general practitioners. The most comprehensive comparison to date is that
of Kwasniewski (2013). He reviewed published true
triaxial strength testing and collated 47 data sets from
2.4 COMPARISONS a variety of rock types, which ranged in UCS values
Many researchers have compared various failure from 7 to 310 MPa. His key observations include:
criteria with laboratory test data. Al-Ajmi and • failure of intact rock is a function of the effective
Zimmerman (2005); Colmenares and Zoback (2002); mean normal stress. A conceptual strength envelope
Costamagna and Bruhns (2007); Kwasniewski (2013); is shown in Figure 2.8 which
Lee et al. (2012); Pan and Hudson (1988); Yoshinaka „„ for brittle fracture is (σ1 + σ3)⁄2,
and Yamabe (1980); You (2009) are just a few that have and
compared various failure criteria with true 3D, or poly- „„ for ductile yielding is (σ1 + σ2 + σ3)⁄3
axial, stress test data. The 3D data sets used in some
of these published comparisons are listed in Table 2.3.

2.4.1 Colmenares & Zoback


Colmenares and Zoback (2002) statistically evaluated
seven criteria against 3D test data (σ1≥σ2≥σ3) in
compressive failure comprising 187 results from five
different rock types. They concluded that using the
Modified Wiebols & Cook criterion (Zhou 1994) or
the Modified Lade criterion (Ewy 1999) gave the best
fit to laboratory test results for rocks with highly σ2 –
dependent behaviour. These were the three extremely
high strength rocks in their database, which had UCS FIGURE 2.8 Conceptual strength envelope for intact rock,
of 300, 375 and 450 MPa. after Sakurai and Serata (2013)

14
2 | Intact Rock Strength

• the effect of σ2 on σ1 at failure is less than 20% for „„ Mogi power-law criterion
low intact strength rocks (defined by Kwasniewski
„„ 3 D Hoek-Brown criterion as proposed by Zhang
as σc < 30 MPa), but up to 50% for very high intact
and Zhu (2007)
strength rocks (σc > 100 MPa)
• the data could be well fitted, as defined by r2 and
standard error of estimate, by either
„„ Mogi linear criterion

TABLE 2.3 Poly-axial data used by various researchers

Source Rock Type Colmen. Al-Ajmi Costam. You Lee Kwasn.


(2002) (2005) (2007) (2009) (2012) (2013)

Haimson & Chang Westerly granite     

Chang & Haimson KTB amphibolite     

Dunham dolomite     

Solenhofen     
limestone

Mizuho trachyte    

Mogi Manazuru andesite  

Yamaguchi marble 

Inada granite 

Orikabe monzonite 

Chichibu schist 

Shirahama      
sandstone

Takahashi & Kiode Yuubari shale     

Horonai sandstone 

Taga limestone 

Michelis Dense marble 

Akai & Mori Izumi sandstone 

Locharbriggs sst 
Smart et al
SSC muddy siltstone 

Slask sandstone 
Kwasniewski
Rozbark sandstone 

Descamps et al Solgnies limestone 

Phra Wihan 
sandstone

Walsri Phu Phan sandstone 

Phu Kadung 
sandstone

Sriapai et al Maha Sarakham sst 

Lee & Haimson SAFOD granodiorite 

15
ESTIMATING ROCK MASS PROPERTIES

2.4.3 Curve fitting and Douglas (2000), Eberhardt (2012) considers that
In their respective comparative analyses, most the closeness of fit at low confinement and in tension
researchers have opted for the lowest mean absolute may be of more concern for rock engineering than at
misfit between the predicted failure stress and the high confinement.
experimental data to determine the best fitting When considering the need for an improved
criterion for their particular data set. However, it can criterion, a major consideration is the inherent
be seen from Table 2.4 that even following this general variability of test data. Sari (2012), for example,
method, the best fitting Hoek-Brown parameters, for conducted a comprehensive program which included
example, do vary slightly between researchers. This 266 tests for one particular rock. The data is plotted
shows that the calculated parameters are not only in two dimensional principal stress space in Figure 2.9
dependent on the rock but are also dependent on the from Bertuzzi (2012). In another example, one of the
curve-fitting technique used. This is in keeping with graphs presented in Colmenares and Zoback (2002)
a conclusion reached by Mostyn and Douglas (2000). is reproduced in Figure 2.10. By inspection of these
Further, the relevance of these comparisons and figures and of the UCS data shown in Figure 2.1, there
the level of fit achieved are also dependent on the is a limit to the amount of “fine-tuning” one can do to
confining stress range selected. Along with Mostyn a criterion to fit test data.

TABLE 2.4 Hoek-Brown parameters from best fitting curves


Hoek-Brown parameters σc [MPa], mi

UCS 2D 3D
Rock Type
[MPa]
Colmenares & Lee et al (2012) Kwasniewski (2013)
Zoback (2002)

KTB amphibolite 165 250, 30.0 168, 46.50 164.7, 31.079

Dunham dolomite 257, 262 400, 8.0 321, 11.58 261.5, 9.653

Solenhofen limestone 310 370, 4.6 299, 8.51 310, 4.442

Mizuho trachyte 100 - 124, 11.11 100, 10.317

Manazuru andesite 140 - 150, 46.49 140, 31.425

Shirahama sandstone 60 65, 18.2 58, 19.10 60, 11.456

Yuubari shale ? 100, 6.5 72, 10.53 68.4, 7.845

Westerly granite 201 - 260, 36.77 201, 32.118

16
2 | Intact Rock Strength

2.5 COMPUTER SIMULATION


It is possible with some constraints to simulate the
laboratory testing of intact rock using particle-based
numerical methods such as a Particle Flow Code
(PFC) model (Potyondy and Cundall 2004). PFC
uses an assembly of spherical particles bonded together
at their contacts to simulate intact core of rock under
various conditions such as UTS, BTS, UCS and TXL.
A number of such laboratory test results are needed
to calibrate the PFC model so that it can be used to
generate a failure envelope (Pye 2009).
The idea is that the PFC model can be used as the
basic block in synthetic rock mass (SRM) modelling.
An SRM is a numerical technique to represent a
jointed rock mass. It does this by combining an
assembly of spherical particles bonded together at their
contacts – a PFC model – with an embedded discrete
fracture network (DFN) (Cundall et al. 2008). The
behaviour of an SRM depends on both the creation
of new fractures through intact material and the slip/
opening of pre-existing joints.
The main advantages of PFC and similar codes is
the relatively simple mathematical formulation and the
codes’ ability to consider material heterogeneity. The
disadvantage is the extensive experimental validation
and calibration required to verify that models are
capturing the observed macroscopic behavior of rock
(Lisjak and Grasselli 2014).
FIGURE 2.9 Sari (2012) figure 5 with the curve generated
following Mostyn and Douglas (2000). This has Hoek- For example, Bahaaddini et al. (2013) investigate
Brown parameters of σc = 70.56 MPa, mi = 7.56 and a = the effect of joint geometry on the strength and
0.590 modulus of a laboratory-scale rock mass specimen
using SRM modelling. One of their conclusions is
that SRM modelling can be calibrated to reasonably
reproduce 2D laboratory UCS and biaxial tests at
low σ2 but the correlation progressively worsens as
σ2 increases. Further, it is understood that PFC does
not yet reproduce the σc / σt ratios observed in hard
rocks, though methodologies of employing ‘clumped
particles’ and ‘actual grain shapes’ to improve this are
being researched (Mas Ivars 2010).
There are further questions as to the application of
PFC to simulate rock mass, in particular the scaling
of the spherical particles to large blocks of intact rock.
This is discussed in Chapter 5.

2.6 CONCLUSIONS
From the discussion presented, it is apparent that
research into intact rock strength is quite mature.
There are many criteria that consider intact rock failure
FIGURE 2.10 Poly-axial test data from Colmenares and but despite their limitations, the Mohr-Coulomb and
Zoback (2002) showing the variability of the test data Hoek-Brown shear strength criteria remain the most
compared with seven intact rock failure criteria
widely used to assess intact rock failure.
The generalised Hoek-Brown criterion has proved
to be a practical method to estimate intact and rock
mass strength (Eberhardt 2012). Though many have

17
ESTIMATING ROCK MASS PROPERTIES

suggested modifications and improvements to the • The Hoek-Brown criterion as modified by Mostyn
criterion since its publication in 1980, e.g. (Benz et al. and Douglas (2000).
2008; Dinc et al. 2011; Ismael et al. 2014; Lee et al. • The Hoek-Brown criterion as extended to 3D by
2012; Mostyn and Douglas 2000; Pan and Hudson Zhang and Zhu (2007).
1988; Saroglou and Tsiambaos 2008; Serrano et al. These criteria have the following characteristics.
2007; Sonmez and Ulusay 2002; Zhang and Zhu
2007), its popularity has remained. Zuo et al. (2015)
• A theoretical basis. It is noted here that while the
Hoek-Brown criterion and hence its derivatives
provide a theoretical basis for the criterion’s parameters. are empirical, the original work did start with the
It is likely that there is a need to have more than one Griffith crack theory. Similarly, the Lade criterion is
criterion to describe all rocks in all stress conditions, as an empirical development from a theoretical basis.
suggested by Descamps and Tshibangu (2007). • A maximum of three parameters. While more
Returning to the principal objective of providing parameters may offer better curve-fitting, such as
robust guidelines for the practicing designer, it is the You criterion (You 2009), they are more difficult
worthwhile comparing the Hoek-Brown criterion with to define particularly with limited test data.
the following four criteria as they too appear to offer • A smooth, curved failure surface in the octahedral
good means of representing intact rock strength. plane and a curved failure surface in planes
• The Lade criterion (Lade 1977). This 3D criterion containing the hydrostatic axis.
was developed from theoretical soil mechanics and • Are capable of having a tensile strength.
laboratory tests. It is a three-parameter criterion In Chapter 8 these criteria are compared with
which has extended into materials which can sustain the Hoek-Brown criterion as well as the Modified
small tensile stresses such as concrete and rock. Wiebols-Cook criterion, which was recommended
• The Christensen criterion (Christensen 1997). This in Colmenares and Zoback (2002) as the
two-parameter, theoretical criterion was developed for criterion that generally best fitted the test data in
true 3D stresses to cater for a wide variety of materials. compressive failure.

18
1 | Introduction

3 Defect
Parameters

3.1 OUTLINE comprises those surfaces; both of which are probably


The preceding chapter looked in detail at intact dependent on scale.
rock strength. In working towards the goal of Direct shear tests can be carried out on laboratory
estimating rock mass properties, this chapter reviews scale samples of defects to measure shear strength. The
properties of defects (also variously called geological results are often corrected for dilation, which is caused
structures, discontinuities, weakness planes, fractures or by the defect’s roughness, and surface area following
simply ‘ joints’). the recommendations made by Hencher and Richards
(1989); (2014). The laboratory test results then need
‘The biggest single difficulty in rock mechanics will to be adjusted in some manner to be suitable for
probably always be the measurement and understanding larger scales, such as those needed for underground
of relevant input concerning joint characteristics’ excavations.
(Barton 1972).
It is worth stating that cohesion for a clean joint can
The main aspects to know about a defect are its be assumed to be zero as any cohesion will be lost at
type (joint, fault, bedding plane, etc.), geometry the onset of movement (Robertson 1970). Hence, the
(orientation, length, spacing, persistence, shape, strength of a defect is largely a function of its friction
thickness, and termination), infill material, hydrology angle (φ).
and strength. Each influences a rock mass’ properties.
These aspects can be observed or measured to varying 3.2.1 Patton
degrees in the field and in the laboratory. Patton (1966) introduced the bi-linear shear strength
While there are good, systematic methods for concept where a defect’s strength is the sum of its
collecting, analysing and presenting defect type and surface’s basic friction angle (φb) and its roughness (i).
to various degrees, defect strength and geometric data, At high normal loads the defect’s asperities are sheared,
many defect properties in fact need to be inferred. For i tends to zero and the defect’s strength becomes φb
example, a defect’s boundary and shape can be inferred (see Figure 3.1). The so-called ‘basic friction’ is the
using “geo-logic” and if lucky, multiple measurements frictional component of shear strength for a planar
of field samples. For instance Zhang and Einstein defect, i.e. independent of any roughness component
(2009) demonstrate planar joints can typically be causing dilation during shear (Hencher and
assumed to be elliptical and of a size that can be Richards 2014).
assessed from trace length data. Stacey and Page (1986) suggest values of φb and i.
Though the following discussion focuses on defect They quote values of φb between 25 and 35° for various
strength and stiffness, it is worth noting that defect rocks and observe that assuming φb = 30° is often
continuity dominates rock mass properties: “the most
important single factor affecting the strength along a
potential failure plane is the continuity of the joints on
that plane” (Robertson 1970). For example, Diederichs
(2003) shows that less than 1% of intact area in the
plane of a defect can provide a distributed load
capacity equal to heavy support systems used in hard
rock mining.

3.2 STRENGTH
The strength of a defect mainly depends on the FIGURE 3.1 Theoretical bi-linear peak shear strength of a
roughness of its contact surfaces and the material that defect from Patton; after Brady and Brown (2005)

19
ESTIMATING ROCK MASS PROPERTIES

adequate for design. These values were also given by however, were researched by Johnston et al. (1993)
Barton (1973). Hencher and Richards (2014) provide and Seidel and Haberfield (2002). On the basis of
values of φb from various work and note typical values theoretical analysis and laboratory testing of clean rock
of 38 – 40° for natural joints through silicate rocks. joints, they concluded that under constant normal
Mineral coatings and infill influence strength. stiffness, the average shear strength of a defect can in
Table 3.1 provides the typical values for i stated fact be approximated by σntan(φb +i). That is, Patton’s
by Stacey and Page (1986). The particular roughness bi-linear equation is applicable to joints under constant
description is scale dependent, and it is inferred from normal force or stiffness.
Robertson (1970) that the descriptions given in Table
3.1 are relative to a scale of 50 mm (core size) to 3.2.2 Barton-Bandis
approximately 300 mm (12” joint blocks). Barton (1973), Barton and Choubey (1977) and
Barton and Bandis (1980) studied the strength and
stiffness of defects and found that a defect’s φ can be
TABLE 3.1 Surface roughness contribution; estimated using:
modified from Stacey and Page
(1986)

Description i
[50 to 300 mm Addition
scale] EQUATION 3.1
to φb

Slickenside or Very smooth, 0 where


Polished reflects light
• JRC is the joint roughness coefficient, which ranges
between 0 and 20 depending on specific roughness
Smooth Roughness +2° profiles
not detected
with finger • JCS is the joint wall compressive strength, which is
equal to σc for unweathered joints but may reduce to
Defined ridges Fine to +6° ≈ σc /4 if the joint walls are weathered (Barton 1976)
(slightly rough) medium
sandpaper
• φr is the residual friction angle for the general case.
For flat, unweathered surfaces Barton and Choubey
feel
(1977) claim φr = φb, i.e. the basic friction angle.
Small steps Medium +10° In their research, Barton & Bandis discovered that
(rough) to coarse increasing block size (L) reduces the values of JRC
sandpaper and JCS (Bandis et al. 1981; Barton and Bandis 1980;
feel Barton and Bandis 1982). They provided the following
relationships where the subscripts 0 refers to laboratory
Very rough Very well +14° and n to in situ scales.
defined
ridges, steps

An evaluation of eight large slope failures by


McMahon (1983) led him to conclude that the defect’s
friction angle (φ) for slope design can be estimated as
the sum of the laboratory basic friction angle φb and the
“relatively large scale” roughness angle: the “relatively
large scale” meaning at least 2% of the potential failure
path (McMahon 1985). So for a 5 m to 15 m long The estimate of JRC and JCS at a scale relevant to
failure, such that might occur around an underground the problem is important. Barton and Bandis (1980)
excavation, the term “relatively large scale” as defined advocate adopting tilt tests and carrying out Schmidt
by McMahon may mean 100 to 300 mm. This is rebound hammer tests on ‘natural’ blocks to calculate
in keeping with Patton’s bi-linear approach and the φr , JRC and JCS at the scale of interest for the problem.
descriptions of roughness given by Stacey and Page Of course, there are questions as to what is and how
(1986) in Table 3 1. to sample a ‘natural’ block at the ‘scale of interest’
for testing.
McMahon (1983) noted that his conclusion was for
defects under constant normal force and may not be The scale dependency of JRC and JCS has also been
applicable to defects sheared under constant normal shown by others. For example, Kveldsvik et al. (2007),
stiffness. Joints under constant normal stiffness who went on to show that the assumed value of JRC has

20
3 | Defect Parameters

FIGURE 3.2 Plot of JRCn versus JRC0 for joint lengths


varying from 1 m to 100 m (Mostyn et al. 1997). The two
example points are for a 100 m long joint. If JRC0 = 6 the
Barton & Bandis equation implies JRCn = 2.5. A JRC0 = 20
however, implies a JRCn = 1.5. That is two joints identical
except small scale roughness, would have reversed larger
scale roughness. Clearly this is wrong. Mostyn et al. (1997)
suggests that the equation may only be appropriate for joint
lengths in the region of 1 m. FIGURE 3.3 Relationship between peak dilation angle
and Barton & Choubey empirical equation. Reproduced
from Hencher and Richards (2014) who re-drew the data
presented in Barton and Choubey (1977).

the greatest effect on the calculated strength. Mostyn the methodology employed by Bandis et al. (1981)
et al. (1997) show that the relationships linking JCS, had ‘severe limitations for accurately and consistently
JRC and L are not suitable for defect lengths greater simulating shear behaviour of most rocks’.
than about 1 m (Figure 3.2). While this is obviously Prediction of friction angle from JRC using
important for rock slopes it is noted that a scale of Equation 3.1 is prone to considerable scatter as evident
“about 1 m” is probably appropriate for underground from Figure 3.3 (Hencher and Richards 2014).
excavation design. Jang and Jang (2014) found that it is better to use φr
Researchers have noted other limitations to the and that it is not the same, and does not necessarily
method. correlate with φb.
Krahn and Morgenstern (1979) and Reeves The French tunnelling association (AFTES 2003)
(1985) are of the opinion that φr and φb are not notes that Equation 3.1 is limited to defects with thin
material parameters and can be reduced to very or no infill. The association also notes that at very low
low values by surface preparation, for example normal stresses it gives unrealistic values of peak shear
grinding and polishing, taking into consideration the strength. This latter aspect is surprising as Barton
dilation induced by the roughness, i. Reeves (1985) and Bandis (1980) suggested method of estimating
further highlighted the following deficiencies of JRC and JCS is using a tilt test which has very low
Barton’s equation. normal stresses.
• The reliability of push or tilt tests which usually
involve very low normal stresses in estimating JRC 3.2.3 Summary
and JCS Estimating a defect’s strength requires some estimate
• A lthough φb for a diamond saw cut may be a of the basic friction angle (φb) which based on the
reproducible “index” for a particular rock type critique of literature presented, is ideally obtained
and particular saw blade, it is not a fundamental from corrected direct shear tests as recommended by
physical property of the material. That is, φb is scale Hencher and Richards (1989); (2014). The defect’s
and test dependent. shape or roughness at a scale commensurate with the
• Both JRC and JCS are found to be strongly scale particular engineering problem in question is needed
dependent parameters. so that φb can be factored to a design shear strength,
• Since JCS and φb tend to be estimated rather than φ. A typical scale of 1 to 10 m is suggested to be
measured, JRC values are therefore inevitably suitable for most underground excavation design.
somewhat subjective. The suggestions made in Table 3.1 regarding surface
Hencher et al. (1993) also showed φb to be variable roughness contribution (i) are recommended.
but questioned the scale effects, suggesting that An alternative way is to use Barton’s method
the differences could be reduced if the results are shown in Equation 3.1 to estimate φ, appreciating the
corrected for dilatancy as recommended by Hencher method’s limitations and that the solution will give
and Richards (1989). They went further, considering a range.

21
ESTIMATING ROCK MASS PROPERTIES

3.3 STIFFNESS
The stiffness of a defect, in both the normal and shear
directions, is a very difficult parameter to measure.
Hudson and Harrison (1997) discussed the behaviour
of defects under load and commented that defect Where Ef If is the foundation stiffness. The value of
stiffness is unlikely to be constant as it varies with the term in square brackets is approximately 1.0 (from
load, as can be seen in Figure 3.4 . chapter 7.5 in Ou (2006)); certainly for a clean, tight
defect, i.e. where there is one material and one modulus
to consider. This equation therefore simplifies to:

Noting that the normal stiffness can be thought of


as the product of subgrade reaction and an area, i.e.
kn = Cr A and A = B over unit length, then combining
these equations gives:

which suggests kn (MPa/m) ≥ Ei (MPa) per m.

3.3.2 Defects with thick infill


For infilled defects the relationship developed by
Sovinc (1969) for rectangular loading on a finite
layer as published in Poulos and Davis (1974) and
FIGURE 3.4 Theoretical behaviour of a defect under reproduced herein as Figure 3.5 can be used. Noting
compression, tension and shear loading (Hudson and that the thickness of the defect (dimensioned as h in
Harrison 1997) Figure 3.5) is typically small compared to the loaded
area (L•B), the influence factor (β) can be approximated
A theoretical approach can be adopted to estimate as the slope of the initial part of the curve in Figure
normal stiffness, kn, for two trivial but instructive 3.5, which corresponds to a value of h / L → 0. This
problems - tight defects and defects with thick infill. gives β ≈ 0.625h/L. Substituting this into the equation
shown in Figure 3.5 for displacement ρz = βpL/E,
3.3.1 Tight defects gives ρz= 0.625hp/E, which re-arranged leads to the
Consider a perfectly tight, perfectly smooth defect approximation of stiffness, kn = p/ρz = 1.6Ej /h where Ej
such that normal compressive loads would behave as if is the defect infill modulus.
the defect was not there. Such a defect effectively does For example, the normal stiffness of a defect with
not have a normal stiffness. This has been confirmed 25 mm soft clay infill (say su = 25 kPa, Ej = 5 MPa)
by researchers, e.g. Bandis et al. (1983); Barton (1972); is kn = 1.6 x 5 MPa / 0.025 = 320 MPa/m. This value
Goodman (1974), who carried out tests on joints and appears to be in keeping with the results published by
found that when joints are nearly or fully closed under Kulhawy (1978).
normal loads, further increases in normal load is Further, the elastic isotropic relationship between
transferred to the solid rock above and below the joint. shear (ks) and normal (kn) stiffness is ks = kn / [2(1+v)],
However, many numerical modelling codes require cf. G = E / [2(1+v)] and Goodman (1968), which
defect stiffness values. It is suggested that the idealised suggests that ks should be 0.33 to 0.5 times kn.
concept of equivalent subgrade reaction (Cr ) can be
employed to estimate a numerical model minimum kn 3.3.3 General defects
for clean, tight defects. Deriving Cr in terms of modulus Unfortunately estimating the stiffness of defects that
from standard elastic solutions for foundations also are between the perfectly tight, perfectly smooth defect
requires the dimensions of the foundation, B. The and the thick infilled defect is not as straightforward.
following equation, for example, is from Vesic (1961) The following are examples of what practitioners have
for an infinitely long and B wide beam as reported in published, often with little substantiating data, as
Chapter 5 of Kulhawy and Mayne (1990). values for normal and shear stiffness values.

22
3 | Defect Parameters

FIGURE 3.5 Influence factor, β, as a function of layer


thickness (h) and loading area (L, B) (from Poulos and Davis
(1974), section 7.8)

• Goodman (1968); (1974) found that defect stiffness FIGURE 3.6 Experimental evidence for scale effect on
varies with load. He suggested a simple method of defect shear stiffness; from Bandis et al. (1983). The data
suggests shear stiffness x length sheared (ks•L) in the range
estimating the initial shear stiffness as ks= σn tan(30
10 to 10,000 MPa, with the majority within 100 to 1000
to 40°) ⁄ δpeak where the peak shear displacement, MPa, for joints in rock.
δpeak is typically in the range 0.3 to 5 mm, average of
2 mm.
• Barton (1972) too found that defect stiffness varies • Bandis (1993) presented an equation for estimating
with load and is scale dependent; the larger the ks from the Barton shear strength equation:
specimen the smaller the shear stiffness.
• Kulhawy (1975) based on a literature search gave
the typical range of ks / kn = 0.04 to 1.2.
• Kulhawy (1978) reported results from various site
tests and showed typical values of kn = 1,000 to EQUATION 3.2
35,000 MPa/m and ks = 100 to 10,000 MPa/m for
joints; and kn = 250 to 7,000 MPa/m and ks = 20 to
2,000 MPa/m for joints with clay. He noted that • Flores and Karzulovic (2002) collated the then
the intact rock modulus to defect normal stiffness available published information regarding joint
ratio, Ei/kn, varied between 0.2 - 4.2 m. stiffness, which is summarised in Table 3.2.
They also provided the equation from Barton (1972)
• Bandis et al. (1983) carried out extensive tests on to estimate the normal stiffness for a simple rock mass
real joints and found for unweathered, high to
very high strength sandstone, siltstone, limestone, containing one defect set, if the intact modulus (Ei),
dolerite and slate that kn = 30,000-270,000 MPa/m rock mass modulus (E m) and defect spacing (s) are
and ks = 10,000-30,000 MPa/m. The laboratory tests known, viz:
showed that a ratio of ks / kn = 0.05 - 0.10 could be
used for normal stresses (σn) greater than about 0.3
MPa and a ratio of 0.10 for σn > 1MPa. They found
stiffness not only varies with load but also with the EQUATION 3.3
defect’s contact surface strength and roughness; and
was inversely proportional to the length of defect
that is sheared (L). Figure 3.6 suggests:
„„ k s•L ≈ a constant. For a given defect, increasing
L proportionately decreases ks. The product k s•L
effectively is in the range of 10 to 10,000 MPa,
and typically between 100 to 1,000 MPa, for
clean rock joints.
„„ ks ∝ normal stress. For a given defect of sheared
length L, increasing σn proportionately increases
ks . Because of the relationship between ks and kn ,
k s•L also increases, presumably to the limit of Ei.

23
ESTIMATING ROCK MASS PROPERTIES

TABLE 3.2 Examples of joint stiffness values joints in weak limestone and marl, which they
from Flores and Karzulovic (2002) believed addressed the variability of the rock mass.
Defect Characteristic kn ks Bandis was one of the authors of the Sharp et al.
[MPa/m] [MPa/m] (2008) paper.
• Mas Ivars (2010) adopts values of kn = 150,000-
250,000 MPa/m and ks = 20,000-30,000 MPa/m, in
Soft infill < 10,000 < 1000
FLAC modelling of joints in volcanic rock masses at
Fault with 300 to 1500 5 5 the Palabora and Northparkes underground mines.
mm thick clay gouge
The values were based on the observed roughness
Defect with rock flour 800 800 and hardness of joints.
infill
Moderately strong 10,000 - < 10,000
• Oliveira (2014) summarises a limited number of
normal loading tests on Hawkesbury Sandstone
rock walls 50,000
and Ashfield Shale that indicate a kn = 4000 MPa/m
Sandstone 7000 - 500 - 4500 for both rock masses under low normal stresses of
25,000
σn ≤ 0.4 MPa.
Limestone - 300 - 5000
Strong rock walls 50,000 - < 50,000 3.3.4 Summary
200,000
The stiffness of three types of defects is considered.
Quartzite 10,000 - 1400 -
30,000 9000 • For perfectly tight, perfectly smooth defects such
that normal compressive loads would behave as
Granite 50,000 - 1000 - if the defect was not there, the minimum normal
635,000 1600
stiffness kn (MPa/m) can be approximated by the
Granite shear 2000 - - rock’s intact modulus Ei (MPa) per m.
266,000
• For infilled defects the relationship shown in Figure
3.5 can be used to estimate normal stiffness as kn =
1.6Ej/h where Ej is the defect infill modulus and h is
• The mass modulus of a Hawkesbury Sandstone
the defect thickness
foundation was measured by Clarke and Pells
(2004) to be between 800 and 1000 MPa with • t is suggested that these bounds loosely define the
I

the corresponding intact rock modulus inferred range of kn for other defects:
to be 3200 to 4800 MPa. Six bedding planes at
an average spacing of 1.25 m, were encountered in
the 9 m deep foundation zone that was tested. The
bedding planes were rock and clay infilled. It is • The published range of defect stiffness is large; kn=
estimated that the normal stiffness of these bedding 10,000 to 250,000 MPa/m. However, for clean
planes to be approximately 1000 MPa/m based on defects, the appreciation that k s•L ≈ a constant
Equation 3.3. (Figure 3.6) suggests that values for k s•L of between
• Badelow et al. (2005) adopt kn = 10,000 MPa/m 100 to 1000 MPa can be typically used.
and ks = 5,000 MPa/m for bedding and joints in • ratio of ks / kn = 0.10 can be assumed for normal
A

sandstone, and kn = 2,000 MPa/m and ks = 1,000 stresses greater than 1 MPa.
MPa/m for bedding and joints in shale. Perhaps
more importantly though, they found in their
numerical modelling using the then current version 3.4 HAWKESBURY
of the software Phase2, that values of kn and ks had SANDSTONE AND
little influence on the end results. Low values of ks
however did increase computational time. Hence, ASHFIELD SHALE
it is important to check if stiffness is materially The results of laboratory direct shear test results which
important in the software used. were undertaken for three recent tunnelling projects
• “Values for normal and shear stiffnesses for rock in Sydney – the Lane Cove Tunnel, the CBD Metro
joints typically can range from 10 to 100 MPa/m (which was not built), and the Northwest Rail Link,
for joints with soft clay in-filling, to over 100 000 comprise 76 tests of Hawkesbury Sandstone defects
MPa/m for tight joints in granite and basalt”, (Itasca and 97 tests of Ashfield Shale defects. The defects are
2008). These values are based on published data either bedding planes or partings, joints or shears.
from Kulhawy (1975), Bandis et al. (1983) and The results largely represent planar defects. In the
Rosso (1976). case of one dataset, the shear and normal stress values
• Sharp et al. (2008) adopt ks = 500-700 MPa/m were also corrected for dilatancy as recommended by
and φ = 40-55° and φr = 30-32°. for bedding and Hencher and Richards (1989). Hence, the data can

24
3 | Defect Parameters

be used to estimate the basic friction (φb), i.e. ”the


frictional component of shear strength for a planar or
effectively planar discontinuity i.e. independent of any
roughness component causing dilation during shear”
(Hencher and Richards 2014).
The results are interpreted for shear strength and
stiffness.
• The peak shear stress (τ) recorded for the normal
stress (σn) applied during each of the direct shear
tests was selected.
• e shear stiffness (ks) was estimated as the gradient
Th

of the secant line plotted on the elastic portion of
the shear stress versus shear displacement graph.

3.4.1 Shear strength FIGURE 3.8 Laboratory direct shear tests results for
bedding and joints in Ashfield Shale.
3.4.1.1 Basic friction angle
Figure 3.7 is a plot of shear stress against normal stress
3.4.1.2 Design friction angle
for the defects in Hawkesbury Sandstone. The similar
plot for Ashfield Shale is shown in Figure 3.8. The defect’s shape or roughness at a scale of 1 to 10
m is suggested to be suitable for most underground
Straight-line conservative estimates of the mean excavation design so that φb can be factored to a design
peak basic friction angle (φb) are shown in these figures, shear strength, φ.
which indicate that within:
Using the examples of bedding planes within the
• The Hawkesbury Sandstone, φb = 30° for bedding Hawkesbury Sandstone, most are logged as “planar,
planes, φb = 40° for joints and φb = 18° for shears /
rough” at the drill core scale (see data which is presented
clay seams
in Chapter 9.2). This can be extended to “smooth
• The Ashfield Shale, φb = 25° for bedding planes, φb to slightly rough” at the 1 to 10 m scale. Using the
= 28° for joints and φb = 15° for shears / clay seams. suggestions made in Table 3.1, the surface roughness
contribution (i) is expected to be in the range of 2 to
6°. This implies a design defect friction angle of 30° +
2 to 6°, i.e. φ = 32 to 36°.
An alternative way to estimate φ is to follow Barton’s
method of Equation 3.1, though as stated previously
the basic friction angle estimated from the direct shear
tests does not necessarily equate to φr in this empirical
correlation (Hencher and Richards 2014; Jang and Jang
2014). For bedding planes in Hawkesbury Sandstone:
•  JRC 0 = 6 to 8 (roughness profile 5 from Barton
and Choubey (1977) and therefore at the 1 m
scale JRCn ≈ 5. [The author is unaware of any tilt
tests carried out on Hawkesbury Sandstone that
according to Barton and Bandis (1980) would
remove any scale dependency.]
•  JCS ≈ UCS ≈ 30 MPa (see Figure 2.1). Any error
caused by incorrectly estimating JCS will be
FIGURE 3.7 Laboratory direct shear tests results for bedding small due to the log formulation (Barton and
and joints in Hawkesbury Sandstone. Bandis 1980).
Therefore, the friction angle of a joint at 50 m depth,
say equivalent to σn = (0.5 to 4) x σv (the multiple of
the vertical stress is to take into consideration the
concentration around a tunnel opening) is: (0.5 to 4) x
50 x 0.025 ≈ 0.6 to 5 MPa, typically 1.25 MPa. That
is, approximately
φ =5log10 + 30=34 to 38°,
typically 37°.

25
ESTIMATING ROCK MASS PROPERTIES

Similarly, for bedding planes within the Ashfield Figure 3.9 and Figure 3.10 indicate:
Shale, which are typically logged as “planar, smooth to • The shear stiffness is dependent on the applied
rough” at the drill core scale (Tables 8.4 to 8.6). They normal stress: shear stiffness increases with
can be taken as “smooth” at the 1 to 10 m scale, which increasing normal stress in the range tested.
implies an i of approximately 2° (Table 3.1), and hence • The shear stiffness of clean bedding planes and
a φ = 25 + 2 = 27°. joints within either the Hawkesbury Sandstone or
Ashfield Shale for normal stresses up to 1 MPa, is
3.4.2 Stiffness generally in the range 300 to 1000 MPa/m.
The shear stiffness inferred from the direct shear tests • There is an indication that the shear stiffness of
is shown in Figure 3.9 for Hawkesbury Sandstone and joints is greater than that of bedding planes.
in Figure 3.10 for Ashfield Shale. The shear stiffness
was taken as the gradient of the secant line plotted
• The shear stiffness of shears and clay infilled

defects either in the Hawkesbury Sandstone or
on the elastic portion of the shear stress versus shear Ashfield Shale for normal stresses up to 1 MPa, is
displacement graphs. This is in keeping with the generally in the range 100 to 400 MPa/m.
approach taken by Bandis et al. (1981).

FIGURE 3.9 Inferred shear stiffness from the direct shear tests results for bedding and joints within Hawkesbury Sandstone.
The shaded region highlights the typical values obtained.

FIGURE 3.10 Inferred shear stiffness from the direct shear tests results for bedding and joints within Ashfield Shale.
The shaded region highlights the typical values obtained.

26
3 | Defect Parameters

Adopting the approximate ratio of ks / kn = 0.10, then 3.4.3 Sheared length


as ks ≈ 300 to 1000 MPa/m for clean bedding planes The maximum shear strength in the direct shear tests
and joints within either the Hawkesbury Sandstone or was obtained with small shear displacements, all less
Ashfield Shale, the implied normal stiffness kn ≈ 3000 than 2 mm and typically 0.2 to 0.5 mm. This is typical
to 10,000 MPa/m. This is lower than that predicted for laboratory direct shear tests, see Bandis et al. (1981);
using the Bandis equation (Equation 3.2) for a Hencher and Richards (2014) for examples.
bedding plane within Hawkesbury Sandstone at 50 m
The length of defect varied between 40 and 80 mm
depth, viz:
depending on drill core size and the relative angle of
the defect to the core axis. The relationship between
shear stiffness and defect length is presented in Figure
3.12, which is in keeping with the laboratory data of
Bandis et al. (1981), see Figure 3.13.
The product of shear stiffness and the initial length
of the defect being sheared (k s•L) is shown Figure 3.14
for clean bedding and joints, irregular bedding and
That is ks = 2700 MPa/m with the corresponding joints and shears / clay infill, respectively. The figures
kn ≈ 27,000 MPa/m. show k s•L increases from shears / clay infill (typically
However, it is in keeping with the value obtained less than 30 MPa) to clean defects (typically 10 to 60
using the simple method suggested by Goodman MPa) and to irregular defects (typically 30 to 70 MPa,
(1968) with φ = 35° for a bedding plane within the with some greater than 100 MPa).
Hawkesbury Sandstone (the range was found to be Limited correlation between shear stiffness and
32 to 36° in the previous section) gives a value for defect length is observable over the scale tested. This
the initial stiffness of ks=1.25 tan 35° ⁄ 0.001, i.e. ks is generally in keeping with the typical range inferred
= 875 MPa/m and by inference, kn = 8750 MPa/m from Figure 3.6.
(ks / kn = 0.10).
It is noted from the observed behaviour of test
samples, that the peak stiffness may be greater than
3.5 CONCLUSIONS
this range (see Figure 3.4). This chapter focused on defect strength and stiffness,
though it is worth repeating that it is defect continuity
Theoretically, Equation 3.3 could be used as an and termination that dominate the large-scale behaviour
alternative means to estimate the stiffness of bedding of rock masses. These latter two defect properties can
planes. For example, the Hawkesbury Sandstone could be inferred from ‘geo-logic’ and measured to varying
be argued to approximate a rock mass with one defect degrees in the field.
set, i.e. sub-horizontal bedding.
Noting the typical values of Ei = 10,000 – 20,000 3.5.1 Strength
MPa and E m = 1,500 – 3,000 MPa for Hawkesbury The following approach for estimating a defect’s
Sandstone and that bedding is typically 2 – 5 m thick, strength is recommended.
the calculations using Equation 3.3 result in a broad
1 Estimate residual or basic friction angle (φr or φb)
range for kn from 300 to 10,000 MPa/m. The most
from corrected direct shear tests as recommended by
likely combinations of Ei, E m and s result in typical
(Hencher and Richards 1989).
values for kn of between 750 to 2500 MPa/m (Figure
2 Add the defect’s shape or roughness at a scale
3.11). Interestingly, this is about an order of magnitude
lower than the test data suggests. commensurate with the particular engineering

FIGURE 3.11 Normal stiffness kn as function of bedding spacing for various intact and rock mass modulus using Equation
3.3. The most likely combinations are shown as solid lines.

27
ESTIMATING ROCK MASS PROPERTIES

FIGURE 3.12 Shear stiffness versus defect length FIGURE 3.13 The data from Figure 3.12 plotted over
Figure 3.6.

FIGURE 3.14 The product ks•L for clean bedding and clean joints (top); irregular bedding and irregular joints (middle); and
shear / clay infill (bottom).

28
3 | Defect Parameters

problem in question. The defect’s shape or • For infilled defects, kn = 1.6Ej/h where Ej is the
roughness at a scale of 1 to 10 m is suggested to be defect infill modulus and h is the defect thickness
suitable for most underground excavation design. • It is suggested that these bounds define the range of
The suggestions made in Table 3.1 regarding surface kn for other defects, i.e.:
roughness contribution (i) are recommended.
An alternative approach is to use Barton’s Equation
3.1 to estimate φ, appreciating the method’s limitations
and that a range of solutions is possible.

3.5.2 Stiffness Defect stiffness can be further estimated noting


It seems that the knowledge of defect stiffness is at the following.
a level that practitioners can readily include it into • The published range of defect stiffness covers just
numerical analyses for underground excavation. The over an order of magnitude; kn = 10,000 to 250,000
data supporting defect stiffness however appears to MPa/m. However, for clean defects, the fact that
be trailing this analytical ability. Nevertheless, the ks•L ≈ a constant (Figure 3.6) suggests that the
stiffness of three types of defects has been considered. range of between 100 to 1000 MPa can be typically
The following approach is recommended. used for values for ks•L .
• The stiffness of perfectly tight, perfectly smooth • A ratio of ks / kn = 0.05 - 0.10 can be assumed for
defects. While a trivial case, this may be required normal stresses greater than 0.3 MPa with that ratio
for analyses such as numerical modelling. For these becoming 0.10 for normal stresses greater than 1
defects, the normal stiffness assigned to the defect is MPa (Bandis et al. 1983).
the equivalent modulus of the rock, i.e. kn (MPa/m)
≈ Ei (MPa) per m.

29
ESTIMATING ROCK MASS PROPERTIES

30
1 | Introduction

4 Rock Mass
Classification

4.1 OUTLINE Rather, in this chapter the two most widely used
There are experienced rock engineering practitioners classification systems for tunnel design – Q, (Barton
who, cognisant of the fact that most failures occur et al. 1974; Barton and Grimstad 2014; Grimstad
due to particular features and not average conditions, and Barton 1993) and GSI (Hoek 1994; Hoek and
consider that geology cannot be averaged and Brown 1997; Marinos and Hoek 2000) – are discussed
quantified into a classification system. However, given specifically in relation to estimating rock mass
that the strength of a rock mass is some combination properties. RMR (Bieniawski 1989) is also used for
of intact rock and defects and that a representative tunnel design, though it is not explicitly discussed as it
volume of a rock mass cannot realistically be tested, is included in the origins of GSI.
some form of empiricism is required. Classification is
often the approach.
It is not the intention to compare the numerous rock 4.2 Q
mass classification systems, which include those in the Barton et al. (1974) in their seminal paper developed
following list. The interested reader is referred to Palmström a correlation between the permanent support installed
(1995) and Edelbro (2003) for such comparisons. in underground excavations with a parameter they
• Terzaghi rock load termed rock mass quality, Q, which comprises:
• Rock Quality Designation (RQD) “the RQD index, the number of joint sets (Jn), the
• Rock Structure Rating roughness of the weakest joints (Jr), the degree of alteration
• Rock Mass Rating (RMR) or filling along the weakest joints (Ja), and two further
• Mining RMR parameters which account for the rock load (SRF) and
water inflow (Jw).”
• Slope RMR
• Rock Mass Index, Q While the terms “the weakest joints” and “ least
• Geological Strength Index (GSI) favourable joint set” were both used in the 1974 paper,
• Rock Mass Index (RMi). the latter term is now solely used, e.g. Barton and
Grimstad (2014); Grimstad and Barton (1993).

FIGURE 4.1 The original data set of some 200 case records upon which the Q system is based (Barton et al. 1974). Palmström
and Broch (2006) identify the Q system as working well for Very Poor to Good quality rock masses in excavations of less than
30m spans, i.e. the shaded portion.

31
ESTIMATING ROCK MASS PROPERTIES

The original work was based on 212 cases. Ninety state that ‘they were mainly from main road tunnels in
seven of the cases came from mapping by Cecil of 18 Norway’. Barton and Grimstad (2014) also report that
projects – 15 in Sweden and three in Norway – mainly an update in 2002-2003 added a further 800 cases but
in granites and metamorphic rocks with high horizontal no details are provided. Graphs showing the scatter in
stress (Cecil 1970), see Table 4.1. Figure 4.1 presents the case data are presented in some publications (e.g.
the original Q log-log chart and the main cluster of case Figure 4.2).
records as defined by Palmström and Broch (2006). The relationship between the Q, excavation span
In 1993, a further 1050 cases were included in the and the support required is shown in Figure 4.3.
Q-system (Grimstad and Barton 1993). Although The Excavation Support Ratio (ESR) for various
there is no detail in relation to the types of rock masses underground excavations is listed in Table 4.2.
of these 1050 cases, Barton and Grimstad (2004)

TABLE 4.1 Case records of rock types used to develop the Q system (Barton et al. (1974), from Cecil (1970))

Igneous 66 Metamorphic 97 Sedimentary 19

Basalt 1 Amphibolite 8 Chalk 1

Diabase 4 Anorthosite (meta-) 1 Limestone 3

Diorite 2 Arkose 1 Marly Limestone 1

Granodiorite 1 Arkose (meta-) 3 Mudstone 1

Quartzdiorite 1 Claystone (meta-) 2 Calcareous Mudstone 1

Dolerite 1 Dolomite 1 Sandstone 4

Gabbro 2 Gneiss 14 Shale 2

Granite 46 Biotite Gneiss 1 Clay Shale 2

Aplitic Granite 1 Granitic Gneiss 4 Siltstone 2

Monzonitic Granite 1 Schistose Gneiss 2 Marl 1

Quartz Monzonite 2 Greywacke 1 Opalinus Clay 1

Quartz Porphyry 2 Greenstone 1

Tuff 2 Schistose meta Greywacke 1

Quartz Hornblende 1

Leptite 11

Marble 1

Mylonite 4

Pegmatite 2

Syenite 1

Phyllite 1

Quartzite 13

Schist 17

Biotite Schist 1

Mica Schist 2

Limestone Schist 1

Sparagmite 2

32
4 | Rock Mass Classification

FIGURE 4.2 The bolt spacing used in case histories (Grimstad and Barton 1993) varies more than what the Q system charts
imply. Graph reproduced from Palmström and Broch (2006).

FIGURE 4.3 Q-chart which shows the relationship between rock mass quality Q, excavation span and recommended support
(Grimstad 2007)

33
ESTIMATING ROCK MASS PROPERTIES

TABLE 4.2 The ESR values recommended in Barton and Grimstad (2014)
Type of Excavation ESR

A Temporary mine openings, etc. ca 2 to 5

B Permanent mine openings, water tunnels for hydropower (exclude high pressure 1.6 to 2.0
penstocks), pilot tunnels, drifts and headings for large openings, surge chambers

C Storage caverns 1.2 to 1.3

Water treatment plants, minor road and railway tunnels, access tunnels 0.9 to 1.1

D Power stations, major road and railway tunnels, civil defence chambers, portals, 0.5 to 0.8
intersections

E Underground nuclear power stations, railway stations, sports and public facilities, factories,
major gas pipeline tunnels

The Q-system does not directly provide estimates of rock mass strength. However, the ratio Jr / Ja is linked to a
frictional strength as indicated in Figure 4.4.

FIGURE 4.4 Approximate relationship between Jr / Ja and apparent ‘shear strength’ (expressed as a friction angle) (Barton
and Grimstad 2014)

34
4 | Rock Mass Classification

Further, the constituent parameters of Q have been In its original formulation GSI ranged from “about
correlated to the Barton-Bandis shear strength criterion 10, for extremely poor rock masses, to 100 for intact rock”
for particular conditions. To recap from Chapter 3, the (Hoek 1994) and was calculated on the basis of the
Barton-Bandis shear strength criterion is: 1976 formulation of RMR (Bieniawski 1976) as:

• GSI = RMR76 for RMR76 > 18 or

For the particular case of clay-filled defects, the • 


GSI = 9Log eQ´+44 for RMR76 <18 where
Q-system parameters Jr and Ja, specifically their ratio Q´=RQD/Jn x Jr /Ja
Jr / Ja, can be used in place of JCSn / σn (Barton 1999).
More generally, Figure 4.5 shows the approximate If the 1989 version of RMR is used as recommended
relationship proposed by Bandis (1993) between the by Bieniawski (1989) then the calculations are:
Q-system parameter Jr and JRC for 20 cm and 100 cm
long defects. So, the Q-system can be used to estimate
a friction angle for some specific cases. • GSI = RMR89 – 5 for RMR89 > 23 or

• GSI = 9Log eQ´+44 for RMR89 < 23

The parameters that make up RMR are listed in


Table 4.3. The range of three parameters changed
between RMR76 and RMR89. The parameter JCond89
is calculated according to its components as given in
Table 4.4.
GSI is therefore a trimmed version of RMR in that
it used only the first four parameters from Table 4.3
assigning the dry score of 10 or 15 to ‘Groundwater’
for RMR76 and RMR89, respectively, and the ‘very
favourable’ score of 0 for ‘Orientation’, to avoid ‘double
counting’ (Hoek 1994). So, the original GSI was a
number based on four parameters; intact rock strength,
RQD, spacing of discontinuities and JCond89 .
Hoek (2007) found the correlation between GSI
and RMR to be problematic for poor quality rock
masses and recommended that the GSI be estimated
directly by means of its charts, as discussed in the
next section.

FIGURE 4.5 Approximate relationship between Jr and JRC TABLE 4.3 P


 arameters used in RMR from
for 0.2 and 1 m long defects (Bandis 1993) Bieniawski (1976) and Bieniawski
(1989)
Parameter Rating
4.3 GSI
Hoek (1994) introduced GSI, essentially as an RMR76 RMR89
alternative to RMR, and correlative with Q, as a means 1 Strength of the intact rock 0 – 15 0 – 15
to determine rock mass strengths primarily: material
• to overcome the limitation that RMR does not work 2 Drill core quality (RQD) 3 – 20 3 – 20
for very poor rock masses
3 Spacing of discontinuities 5 – 30 5 – 20
• to avoid double counting the influence of the
groundwater in RMR or the joint water reduction 4 Condition of discontinuities 0 – 35 0 – 30
factor ( Jw) and the stress reduction factor (SRF) (JCond)
in Q
5 Groundwater 0 – 10 0 – 15
• o have a “system based more heavily on fundamental
t

geological observations and less on ‘numbers’ ” 6 Adjustment for -60 – 0 -60 – 0
discontinuity orientation
(Hoek 2002).

35
ESTIMATING ROCK MASS PROPERTIES

TABLE 4.4 Components of JCond89 after Bieniawski (1989)


Length <1m 1–3m 3 – 10 m 10 – 20 m > 20 m
Rating 6 4 2 1 0

Separation None < 0.1 mm 0.1 – 1 mm 1 – 5 mm > 5 mm


Rating 6 5 4 1 0

Roughness Very rough Rough Slightly rough Smooth Slickensided


Rating 6 6 6 1 0

Infilling None Hard < 5mm Hard > 5mm Soft < 5mm Soft > 5mm
Rating 6 4 2 2 0

Weathering Unweathered Slightly Moderately Highly Decomposed


Rating 6 5 3 1 0

4.3.1 Chart
Following on from the seminal work of Hoek and GSI (Figure 4.6). This paved the way for assessing GSI
Brown (1980), in which a table was published that directly from a chart rather than quantifying through
linked a geological description to an RMR value and to RMR. An early chart was published in Hoek and
the Hoek-Brown strength criterion parameters, Hoek Brown (1997) and is reproduced here as Figure 4.7.
(1994) published a chart showing that linkage with

FIGURE 4.6 Chart published in Hoek (1994) that linked a geological description, GSI and the Hoek-Brown criterion

36
4 | Rock Mass Classification

FIGURE 4.7 Original GSI chart from Hoek and Brown (1997)

Significantly, Marinos and Hoek (2000) largely in the original work of Hoek and Brown (1980) as
through work done in Greece on tectonically disturbed reproduced here in Figure 4.8.
and/or weathered rock masses, extended the range It is worth noting that as the problem scale increases
covered by the chart and suggested GSI be estimated further from those shown in Figure 4.8, the rock mass
directly from it. Estimating GSI directly from the may again be dominated by one or multiple large
chart was also the recommendation independently geological structures such as regional faults. That is
reached by Mostyn and Douglas (2000). the ‘intact rock’ is replaced by ‘rock mass’ separated by
Several implications arise from estimating the GSI regional faults or persistent defect sets.
value directly from the chart instead of calculating A third implication is that the interlocking of rock
it through RMR. The first is that though intact rock blocks was introduced.
strength is broadly implied by selecting a particular
chart e.g. sandstone or granite (Marinos and Hoek, There is actually another more pressing issue with
2000), strength is not explicitly included. In contrast, estimating GSI directly from the chart, and that is
intact rock strength is the first parameter required selecting which chart to use.
when estimating RMR. The second is that problem An important development of GSI in the past
scale needs to be considered by the experienced decade, again largely through work done in Greece
practitioner when selecting the vertical axis of the on tectonically disturbed and/or weathered rock
chart. By so doing, one of the RMR / GSI limitations is masses, is the recognition that general categorisation
addressed. This follows on from the notion expounded of rock masses is complex. The first development was a

37
ESTIMATING ROCK MASS PROPERTIES

Carter and Marinos (2014) prepared a matrix


showing the relationships between the nine GSI charts
and the Hoek-Brown strength criterion parameters
of σc and mi for different lithologies. This highlights
the importance of developing the geological model to
estimate GSI.
One of the main issues with the GSI chart(s) is
that details of the calibration carried out to derive
it is lacking. Marinos (2010) states that the GSI was
compared with the temporary support installed.
However, this is at best an indirect comparison given
that GSI is not used to estimate support but rather
to estimate rock mass properties, which in turn are
presumably used in analyses to estimate the rock
mass behaviour, which in turn is used to estimate
rock support.
FIGURE 4.8 Various rock mass problem scales from Hoek In addition to the need for different sets of material
and Brown (1980)
properties and by extension GSI values, to cater for
different problem scales, it is suggested that different
separate GSI chart for schist around Athens (Hoek et numerical methods also require different values. For
al. 1998), followed by a chart for flysch (Marinos and example, it is logical that the GSI value for the same
Hoek 2000). Separate charts were then introduced for rock mass increases as the analysis method and scale of
molasse (Hoek et al. 2005) and ophiolites (Marinos et the problem reduces:
al. 2006). There are nine GSI charts to cover general • It may only be possible to include a few specific
jointed rock, flysch (two charts), molasse (fissile and geological structures in a large numerical model
confined), limestone, ophiolites and gneiss (Marinos of an open pit. Hence, properties to be used in
2010) and metamorphic rock masses (Habimana et continuum analyses e.g. FLAC, Phase2, PLAXIS, are
al. 2002). Each of the individual GSI charts describes likely to be appropriate.
the particular rock mass type in greater detail than the • or a smaller problem, such as tunnel scale or
F

generic description originally provided. However, some multiple bench scale in an open pit, it may be
of the charts seem to slightly extend the applicability of practicable to include several specific geological
the original GSI. structures in the continuum analyses models.
The sketches on the left-hand side of the GSI chart(s) Hence, different material properties to those used
that are drawn to assist allocating the rock mass into in the first problem are likely.
the appropriate row have also slightly changed over • If that problem was analysed using discontinuum
time. As can be seen in Figure 4.9, the sketch for a techniques e.g. UDEC, or ‘jointed’ continuum
‘BLOCKY’ rock mass is subtly different in the various analyses with numerous geological structures explicitly
publications. It can be seen though that the closer joint modelled e.g. jointed Phase2, it may be appropriate
spacing is kept in Hoek et al. (2013). It is also kept in that material properties reflecting the behaviour of a
the program RocLab (3). smaller representative rock mass are used.

FIGURE 4.9 The changing sketches depicting the rock mass structure. From left to right the sketches are from: Hoek (1994),
Hoek et al. (1998), Marinos and Hoek (2000), Marinos et al. (2006), Marinos (2010) and Hoek et al. (2013).

(3)
Version 1.033 build date 8 July 2013 RocScience Inc

38
4 | Rock Mass Classification

The GSI chart’s axes are not necessarily independent. 4.3.2 Quantification
Often the blockiness of the rock mass (vertical axis) is Several authors have suggested methods to quantify
linked to the surface quality of a rock mass (horizontal the GSI chart ’to facilitate use of the system especially by
axis). This is commonly seen as a diagonal trend from inexperienced engineers’ (Cai et al. 2004). The issue of
top left to bottom right when different classes of the inexperience is perennial; one shared by many leading
same rock mass type are overlain on one chart. For rock engineering practitioners, for example Terzaghi
example, the charts published by Marinos and Hoek (1946), Bieniawski (1976), Hoek (1994), Fookes
(2000), Habimana et al. (2002) and Marinos et al. (1997), Stacey (2000), Edelbro et al. (2007), Pells
(2006), who also noted that not all combinations are (2008). Yet industry relies on young geologists and
possible and excluded the top right and bottom left engineers for routine field work. Quantifying the GSI
cells in the GSI chart (Figure 4.10). may provide a means of reducing inadvertent errors and

FIGURE 4.10 Ranges of GSI for various qualities of peridotite-serpentinite rock masses in ophiolitic complexes from Marinos
et al. (2006). The chart shows the typical diagonal trend from top-right to bottom-left for decreasing rock mass quality.

39
ESTIMATING ROCK MASS PROPERTIES

inconsistencies by these inexperienced professionals in volume’ ( Vb) and the horizontal axis ‘joint condition’
classifying a rock mass. ( Jc ) following the RMi classification system (Palmström
Sonmez and Ulusay (1999) suggested linking the 1995). The joint condition factor is defined by the large
vertical axis of the GSI chart to ‘volumetric joint scale ( JW) and small scale ( Js) roughness and alteration
count’ ( Jv) and the horizontal axis to three of the five ( JA), i.e. Jc = JW x Js/JA (see Table 4.5 to Table 4.7). A
components of the RMR’s JCond parameter. similar approach was taken by Russo (2009) but gave
an explicit equation between GSI, Vb and Jc.
Cai et al. (2004) noted that by adding measurable
data, the system becomes less dependent on the Hoek et al. (2013) waded into the argument of
practitioner’s experience. They proposed quantifying quantifying GSI and, seemingly counter to one of
the GSI chart by making the vertical axis ‘block the system’s original purposes vis a vis “a system based
more heavily on fundamental geological observations
and less on ‘numbers’” (Hoek and Marinos 2007), and
TABLE 4.5 Terms to describe large-scale proposed to quantify the GSI. The reason given by
waviness Jw from Cai et al. (2004)
Hoek et al. (2013) for quantifying GSI is the same as
Waviness term JW others in that the GSI chart is being used by
Interlocking 3 inexperienced engineers ‘who are less comfortable with
these qualitative descriptions’.
Stepped 2.5
Hoek et al. (2013) quantify GSI as equal to 1.5 x
Large (>3%) undulation 2 JCond 89 + RQD/2 (Equation 1) and as shown in Figure
Small to moderate (0.3 to 3%) undulation 1.5 4.12, correlate the chart’s vertical axis to RQD/2 and
its horizontal axis to 1.5 x JCond 89. It can be seen that
Planar (<0.3%) 1
the top and bottom rows are removed compared with

TABLE 4.6 Terms to describe small-scale roughness JS from Cai et al. (2004)
Roughness Description Js
Very rough Near vertical steps and ridges occur with interlocking effect on the defect surface 3
Rough Some ridge and side angle are evident; asperities are clearly visible; defect surface feels 2
very abrasive (rougher than sandpaper Grade 30)
Slightly rough Asperities on defects are distinguishable and can be felt (like sandpaper Grade 30-300) 1.5
Smooth Surface appears and feels smooth (smoother than sandpaper Grade 300) 1
Polished Visual evidence of polishing exists. This is often seen in coating of chlorite and talc 0.75
Slickensided Polished and striated surface that results from sliding along a fault surface 0.6-1.5

TABLE 4.7 Terms to describe joint alteration factor JA from Cai et al. (2004)

Term Description JA

Healed or welded joints (UW) Non-softening, impermeable filling (quartz, epidote, etc.) 0.75
Rock wall contact

Fresh rock walls No coating or filling on joint surface, except for staining 1

Alteration of joint wall The joint surface exhibits one class higher alteration than 2
(SW-MW) the rock

Alteration of joint wall (HW) The joint surface exhibits two classes higher alteration than 4
the rock

Sand, silt, calcite etc. Coating of frictional material without clay 3


partial or no contact

rock wall surfaces

Clay, chlorite, talc, etc. Coating of softening and cohesive material 4


Filled joints with

between the

Sand, silt, calcite etc. Filling of frictional material without clay 4

Compacted clay materials Hard filling of softening and cohesive materials 6

Soft clay materials Medium to low over-consolidation of filling 8

Swelling clay materials Filling materials exhibits swelling properties 8-12

40
4 | Rock Mass Classification

FIGURE 4.11 Quartzite rock mass downstream of the Ord Diversion Dam, Western Australia (photo courtesy of Kurt Douglas)

the GSI chart of Figure 4.10, reverting to the original rock mass behaviour will be increasingly controlled by
rock mass structure range in Hoek (1994) and Hoek the large, persistent defects. Further, neither of these
and Brown (1997) but with slightly lower GSI values. ‘blockiness’ measures capture the degree to which rock
It is important to note that the applicable range for a blocks are interlocked, which is an important part of
reliable quantified GSI is limited to RQD ≤ 80. the descriptions in the GSI chart. For example, the
photograph of the rock mass in Figure 4.11 shows a
It is not unexpected that there is correlation between
very well interlocked blocky rock mass, which plots
GSI and the two RMR parameters JCond89 and RQD. between GSI = 50 to 70. However, the rock mass has
‘Groundwater’ and ‘Discontinuity Orientation’ are not essentially RQD = 0.
part of GSI, and there is clearly an overlap between
It is important therefore to re-iterate the
RQD and ‘Spacing of discontinuities’. Hence, the six
recommendation of Cai et al. (2004) and Russo (2009)
parameters that make up RMR (see Table 4.3) are
and implied in Hoek et al. (2013) that the quantified
really reduced to the two in GSI. GSI approach is supplementary to and not necessarily
Defining the vertical axis as equivalent to Jv , Vb or a replacement of the visually assessed GSI.
RQD/2 implicitly re-introduces problem scale though.
The same rock mass generates the same GSI irrespective
of scale. Hoek et al. (2013) caution that using RQD, 4.4 LIMITATIONS
Jv or Vb ‘ limits the defintion of rock structure to the Both Q and GSI use geological and geometric
dimension of the blocks’. Cundall et al. (2008) suggested parameters to arrive at a quantitative value to represent
addressing this shortcoming by expressing the vertical the quality of the rock mass. By the very nature of rock
masses, it is to be expected that a large scatter exists in
axis “as the number of blocks across the scale of interest.”
the data that lies behind the empirical classifications
Also, RQD, Jv or Vb essentially assume that all (see Figure 4.2 for example). This inherent difficulty
defects contribute equally to the rock mass properties. with all classification systems means that correlations
However, as the scale of the problem increases, the and the ensuing results must be used cautiously.

41
ESTIMATING ROCK MASS PROPERTIES

FIGURE 4.12 Revised GSI chart from Hoek et al. (2013)

42
4 | Rock Mass Classification

FIGURE 4.13 Comparison of the range covered by RQD with those of joint spacing, block volume and volumetric joint count;
from Palmström (2005)

Q and GSI use scale-dependent RQD and joint approach, such as Q or GSI, is required to estimate
condition as direct ratings. Hence neither may rock mass properties.
adequately consider the scale of the rock mass problem, It appears that GSI continues to offer good
e.g. Hudson and Harrison (1997), Douglas and Mostyn opportunity to estimate rock mass properties,
(1999). Palmström (2005) considers that Q and RMR despite its inherent limitations. Particularly for most
(and by extension GSI) would be improved if ways of underground excavations the scale-dependency of
capturing block size measurements other than RQD
RQD may not be too significant. Perhaps the GSI
had been used, and recommends volumetric joint
chart-based classification system should be modified to
count see Figure 4.13.
become more objective as recommended by Cai et al.
In addition to the inherent variability of the (2004) and Hoek et al. (2013). This could be achieved
underlying data, the subjectivity of the person using by quantifying:
the classification method also controls to a large part
the results obtained from any classification method • The horizontal axis with defect conditions, as say
(Edelbro et al. 2007). per Cai et al. (2004) or by Hoek et al. (2013)
• The vertical axis by problem scale, say by inclusion
of direct block size measurements, as suggested by
4.5 CONCLUSIONS Palmström (1995) or by a means that relates the
Accepting that a representative volume of a rock mass block size to the problem scale and analysis method
cannot realistically be tested, a rock mass classification as supported by Cundall et al. (2008).

43
ESTIMATING ROCK MASS PROPERTIES

44
5 | Rock Mass Strength Criteria

5 ROCK MASS
STRENGTH CRITERIA

5.1 OUTLINE This empirical criterion is broadly consistent with the


extension strain criterion for intact rock discussed
The preceding chapters critically reviewed intact
in Chapter 2 and with the research outlined below,
rock strength, strength of defects and rock mass
which suggests spalling initiates at stress levels of
classification. In this chapter, the approach and bases
approximately σ1/σc = 0.3 – 0.5.
of the strength criteria most commonly adopted for
rock masses are assessed. These include: • Stacey (1981) developed an extension strain failure
criterion for brittle rocks stating that rock fracture
• Massive brittle rock, of which coal is a particular initiates when strains exceed some critical value
example
which appears to correspond to a stress level of σ1
• Plane of weakness ≈ 0.3σc.
• Barton-Bandis • artin et al. (1999) and Kaiser et al. (2000) found
M

• Hoek-Brown that brittle failure initiates when σ1 ≈ 0.4σc. They
• Cohesion weakening – Friction strengthening also stated that the depth of stress-induced brittle
(CWFS) fracturing can be found by considering only the
• Synthetic Rock Mass (SRM). cohesive strength of the rock mass, i.e. constant
deviatoric stress (σ1 - σ3) equal to approximately
0.3σc to 0.5σc.
5.2 MASSIVE BRITTLE ROCK
Research by Wilson (1971) which is quoted in Hoek
• Kaiser et al. (2000) also suggested the idea of a
spalling limit. Values of σ1/σ3 ≥ 10 to 20 coincide
and Brown (1980) and in Stacey and Page (1986) led with localised zones of tension that promote
to an empirical failure criterion based on the ratio of spalling and failure. Kaiser et al. (2000) elaborated
vertical stress to UCS (σv/σc ) as listed in Table 5.1. on previous work that the depth of failure (df )
The criterion is applicable to massive, brittle rock at a surrounding a tunnel of diameter a driven in brittle
problem scale of 3 to 4 m square tunnels. rock is proportional to the stress ratio σ1/σc, viz:

TABLE 5.1 Massive, brittle rock failure criterion;


from Stacey and Page (1986) quoting
Wilson (1971)
σv /σv Description of Condition When df reaches half the pillar width, Martin et
< 0.2 No particular problems al. (1999) suggest that the pillar must be rated as
unstable with a high potential for complete failure.
0.2 – 0.4 Spalling from surface parallel to σ1;
Heavier support required
• Diederichs et al. (2004) summarised the findings

of several co-researchers and found that failure in
0.4 – 0.5 Major spalling; Heavier support hard rock initiates at stress levels approximated by
required σ1 = (0.3 to 0.5)σc + (1 to 1.5)σ3 (Figure 5.1).
0.5 – 0.67 Very dangerous and difficult to keep
open; Support heavy and costly

> 0.67 Impractical or extremely difficult to


maintain open

45
ESTIMATING ROCK MASS PROPERTIES

or empiricism to identify the threshold of UCS*: the


UCS at which cracks initiate and accumulate.
The empirical relationships are given in Equation
5.2, where a = 0.2 for ‘damage initiation’ and 0.25
for ‘systematic cracking’ and T is the intact tensile
strength. In a follow-up paper, Diederichs et al. (2007)
recommends the ‘systematic cracking’ envelope (i.e. a
= 0.25) to model spalling.

EQUATION 5.2
FIGURE 5.1 Correlation between estimated minimum in
situ strengths and crack initiation; figure from Diederichs et
al. (2004)

• This was later modified slightly to (Diederichs 2007):

EQUATION 5.1

Equation 5.1 is in a similar format to the Mohr-


Coulomb shear strength criterion (Chapter 2.2.1),
which can be written as σ1=σc+kσ1.
Diederichs (2000) working with computer models
and test results of the Lac Du Bonnet granite of
AECL’s URL in Manitoba, Canada, noted that brittle
rock under low confinement near excavations, fails by FIGURE 5.2 Curved strength envelope suggested by
fracturing, spalling and in the extreme, slabbing. He Diederichs (2007)
found that conventional rock mechanics models, such
as Mohr–Coulomb and Hoek–Brown, do not consider Brittleness is not solely a material property as it
this failure process properly. depends on the stress path the rock mass follows before
Diederichs (2000) proposed the damage initiation and after failure (Mas Ivars 2010). For example, a high
spalling limit (DISL) approach to model the strength confining stress makes a rock mass’ post-peak response
of spall-prone rock masses. The DISL strength envelope more ductile. Therefore, the post-peak deformation
is schematically shown in Figure 5.2. It comprises four properties are also key aspects for understanding
zones of rock mass failure.
• Below the ‘‘damage threshold’’ the rock is not
damaged and remains undisturbed.
• Depending upon the confinement when this
threshold is exceeded, micro-crack damage
accumulates, leading either to
„„ Spalling or
„„ Macro-scale shear failure
• Under tensile stresses, the rock mass can unravel.
In a series of papers, Carter et al. (2007); Carter et
al. (2008); Carvalho et al. (2007); Diederichs (2003);
Diederichs (2007); Diederichs et al. (2007); Diederichs
et al. (2004) suggested an empirically-based method
to derive equivalent Hoek-Brown fracture envelopes
for the first part of the DISL, which he and his co-
researchers variously refers to as ‘damage initiation’
and ‘systematic damage’ or ‘systematic cracking’. The FIGURE 5.3 Tensile crack initiation (CI) stress versus UCS
method uses acoustic emission data, radial strain data for various rock types; figure from Hoek and Martin (2014)

46
5 | Rock Mass Strength Criteria

the behaviour of rock masses. However, post-peak Barton and Shen (2016) suggest that UCS/σt = 10
properties are beyond the scope of this book. and v = 0.25 for most rocks, so
Hoek and Martin (2014) noted the consistency of
tensile crack initiation observed in laboratory tests as
evidenced by the graph in Figure 5.3 which suggests
that the crack initiation stress is approximately equal
to 0.45 times the UCS, i.e. σ1 ≈ 0.45σc. Using typical ratios for UCS/σt of between 10 to 20
(Lade 1993) and for v of between 0.2 to 0.3 (Gercek
Using the findings of the above quoted research,
2007), the equation suggests a range of: σtangential =
that the value of UCS* is between 0.3 to 0.5, but
0.2UCS to 0.5UCS.
commonly 0.45, times UCS, Equation 5.2 for
‘systematic cracking’ then yields s ≈ 0.45(1/0.25) = 0.041
and m = 0.041*(UCS/|T|). 5.3 COAL
Barton and Shen (2016) provide an argument that
the crack initiation stress ≈ 0.4 UCS is not surprising. 5.3.1 A brittle rock
They start with the extension strain criterion (Stacey In terms of rock failure, coal can be thought of as a
1981) which from Chapter 2.3.5 is: brittle rock even at moderate confining pressures.
The triaxial data of coal samples from Medhurst
and Brown (1998) is plotted on a principal stress plot
as an example in Figure 5.4 with Mogi’s line depicting
In two-dimensional plane stress, the strain is related
the approximate transition between ductile flow and
to stress as:
brittle fracture. It is noted that Mogi’s line is often
presented as σ1/σ3 = 3.4 (e.g. the software program
RocLab by RocScience) rather than (σ1- σ3 )/σ3 = 3.4
Tensile failure at the wall of an underground as shown in Mogi (1966); (1971) and discussed in
excavation (σradial = 0) will occur when the critical Douglas (2002). Both lines are shown in Figure 5.4,
strain is reached (σradial=- σ1/E) i.e. which indicates that at a laboratory scale, intact coal
plots well within the brittle fracture zone even up to 10
MPa confinements.
The results of compressive testing of various lengths
of 54 mm diameter cylindrical samples of coal showed
a range of behaviours from brittle to ductile strain-

FIGURE 5.4 Left - Coal triaxial test data from Medhurst and Brown (1998). Right - Load versus displacement curves from
laboratory testing of 54 mm diameter cylindrical coal samples. The curves show strain-softening becoming strain-hardening
post-yield behaviour depending on w/h ratio, i.e. confinement; from Das (1986).

47
ESTIMATING ROCK MASS PROPERTIES

softening and ductile strain-hardening with increasing


confinement Figure 5.4 (Das 1986),
The behaviour of coal pillars under compression
also varies from brittle to ductile dependent upon the EQUATION 5.4
degree of confinement as demonstrated with the in situ
tests carried out by Wagner (1974) and by numerical
van Heerden (1975) describes the in situ compressive
modelling (Esterhuizen et al. 2010; Zipf 1999).
testing of 10 large coal samples. Combining his results
with previous testing, he confirmed Bieniawski’s
5.3.2 Coal pillar strength conclusion that beyond a certain critical size, pillar
The process for estimating the rock mass strength strength does not increase for fixed h. van Heerden
of coal pillars has developed along a different path (1975) proposed the relationship shown in Equation
to that for other rock types. This different approach 5.5 where the Constant is the compressive strength of a
and the seminal papers of this research area are briefly 1.5 m3 cube sample of coal.
presented in this section.
5.3.2.1 Development
In 1960 at the Coalbrook Colliery in South Africa,
more than 7000 pillars collapsed, over 4400 coal EQUATION 5.5
pillars failed in a 5-minute period; 437 men were
killed in the collapse. The Coalbrook disaster was the
impetus for research into coal pillar design. Salamon The Constant in Equation 5.4 ≈ 4.2 MPa.
and Munro (1967) sent out a questionnaire to collieries Sheorey et al. (1987) assessed the details of 23
in the Transvaal and Orange Free State of South failed and 20 stable pillars from Indian coal mines
Africa requesting detailed dimensions of workings and concluded that the in situ strength of coal pillars
(stable areas) and areas of pillar collapse. They collated was more affected by the depth of cover (H) than the
125 case studies of which 98 were stable and 27 were results of laboratory tests. They proposed the following
collapsed where the coal seam was the weakest element. equation.
Starting with the work of previous researchers, they
empirically derived Equation 5.3; where the value K
is the strength of a unit 1 ft3 cube of coal which they
found could be taken as 1322 psi. In SI units, the
value of K for a unit 1 m3 cube of coal is 7.17 MPa, e.g.
van Heerden (1975), Galvin et al. (1999) and York et
al. (2000). σcube is the strength of a 25 mm sized cube

EQUATION 5.6

EQUATION 5.3 A slightly truncated database of 14 failed and 14


stable coal pillars from India was back-analysed using a
Hoek-Brown strain-softening criterion by Jaiswal and
Bieniawski (1968) reported the results of insitu tests
Shrivastva (2009). They derived the following equation
on 25 square coal specimens measuring 2 to 5 ft and
for pillar strength by curve-fitting the analysis results.
up to 5 ft in height. He concluded that the “strength
of pillars cannot be expressed by a power function of their
width and height as there exists a critical pillar [cube]
size above which the strength of coal does not change with
an increase in pillar dimensions”. For example, if the Their paper also included average UCS results from
critical size is 5 ft, then a 20 ft cube pillar will have the Sheorey et al. (1987) which ranged between 19 and
same strength as a 5 ft cube pillar. 50 MPa with a mean of 34 MPa. Using the mean, the
The problem with a power function is the equation simplifies to Equation 5.7.
unreasonable implication that pillar strength continues
to increase indefinitely as the size of cubic pillars (w
= h) increases. Hence, Bieniawski offered a linear
relationship in terms of width / height ratio for w/h ≥ 1
and w ≥ 5 ft (1.5 m) (Equation 5.4): EQUATION 5.7

48
5 | Rock Mass Strength Criteria

A large study of UCS tests and coal pillars in the US Their linear form for pillar strength applicable in the
by Mark and Barton (1997) led them to the thought- range w/h ≤ 8 relationship is given in Equation 5.11.
provoking conclusion that “ laboratory testing should
not be used to determine coal strength”. Their tabulated
data of the averages of nearly two thousand UCS tests
from 54 coal seams ranges between 4 and 47 MPa with
EQUATION 5.11
an overall average UCS of 23.4 MPa.
They correctly point out the strong influence that
geological structure has on the rock mass strength of
a coal pillar. Laboratory tests, particularly those of
blocky coals, require a significant amount of fracturing
of intact coal. Pillars contain so many cleats and other
discontinuities that their failure can occur almost
entirely along pre-existing fractures. The laboratory
tests measure a parameter – the intact coal strength –
that is apparently irrelevant to the insitu strength.
Mark and Barton (1997) found a linear relationship
similar to Equation 5.5 between Cp and w/h – the
Mark-Bieniawski formula, Equation 5.8. They
recommended adopting a uniform 6.2 MPa for “coal
strength” for more reliable pillar designs (4).
FIGURE 5.5 Plan view of coal pillar geometry, after Galvin
et al. (1999)

EQUATION 5.8
Though the relationship is said to be applicable
for w/h ≤ 8, only one of the collapsed pillars in the
This was extended to rectangular coal pillars of database has a ratio w/h = 8; the others having w/h
length L by Mark and Chase (1997) as: < 5 (Galvin et al. 1999). Further, there is debate as to
whether the one case with w/h = 8 is valid; see Colwell
(2010) and Seedsman (2012). However, Galvin (2006)
discusses the lack of failure data for pillars where
w/h > 5 and provides the following two reasons.
EQUATION 5.9
• The loading regime necessary to cause failure of
squat pillars is not generated in bord and pillar
Mark and Barton (1997) did not prove that all coal mining layouts. The one collapsed case with w/h
seams actually are the same strength, their work only > 5 was deliberately generated by progressively
showed that laboratory testing was no help in identifying trimming the sides off pillars.
the differences that surely exist (Mark 2006).
• Failure of squat pillars is characterised by strain
Galvin et al. (1999) adding to the work of Salamon softening rather than total collapse. The strain
and Munro (1967), developed a similar empirical softening may go unnoticed. The one collapsed case
relationship for the strength of coal pillars in Australia with w/h > 5 was detected by an extensive array
based on the following data. of monitoring.
• An Australian database of 19 collapsed and 16 The influence of discontinuities, surrounding strata
un-collapsed pillars characteristics and deterioration with time on coal
• A South African database of 42 collapsed and 98 pillar strength was observed by York et al. (2000).
un-collapsed pillars. They attempted to introduce a classification system
They also recognised the influence of rectangular to consider joint frequency, condition and orientation
pillars and recommended calculating effective pillar into coal pillar strength.
width as the harmonic mean according to Equation The South African coal pillar database was expanded
5.10 and pillar geometry as shown in Figure 5.5 where to 54 cases of collapsed pillars by van der Merwe
w1 ≤ w2 . (2003) who recommended separating ‘weak’ and
‘normal’ coal in using the linear relationship between
Cp and w/h (Equation 5.12). van der Merwe (2003)
argued that the linear relationship is appropriate for
EQUATION 5.10 coal pillars larger than 1.5 to 2 m. The constant for
(4)
The pillar design recommended by Mark and Barton (1997) is the Analysis of Retreat Mining Pillar Stability (ARMPS) method

49
ESTIMATING ROCK MASS PROPERTIES

‘normal’ coal is the range 2.8 to 3.5 MPa, with a mean A point that should be appreciated is that the
of 3.15 MPa. formulae given above in Equation 5.3 to Equation 5.13
recommend slightly different Factors of Safety (FoS) to
be used. For instance,
EQUATION 5.12 • Mark and Barton (1997) suggest FoS = 1.5
• Galvin et al. (1999) suggest FoS = 1.85 or 2.68 to be
equivalent to 1 pillar failure in 1,000 or 1,000,000,
The concept of separating different strength coals
respectively
was extended by Salamon et al. (2006) who further
expanded the South African database to 77 collapsed • van der Merwe (2003) suggests FoS = 1.6
and 245 un-collapsed pillars. The data led them to • Salamon et al. (2006) recommend a ‘probability of
conclude that “coal is not homogeneous and using a single survival’ rather than a FoS.
formula for the design of coal pillars in all coalfields in These differing FoS can be thought of as modifying
South Africa is not ideal”. the calculated value of Cp.
The South African database now comprises 86 Another point worth noting is that Seedsman
collapsed and 337 un-collapsed pillars (van der Merwe (2012) concludes that the equation of Galvin et
and Mathey 2013). The w/h for the expanded collapsed al. (1999) (Equation 5.11), and hence presumably
cases in South African database ranges between 0.9 the other empirical pillar formulae, should not be
and 4.45, i.e. all with w/h ≤ 5. Importantly, no pillar extrapolated beyond the empirical database, i.e. w/h
failure has been recorded in South African collieries ≤ 4.8. Seedsman (2012) further opines that if the
with a w/h ratio greater than 4.45. equation is used for w/h > 5 then it probably results in
Salamon et al. (2006) proposed three empirical unnecessary sterilisation of coal.
relationships; a power, a linear and a non-linear form 5.3.2.2 Other
where the parameters - Constant, α, β, ω and r - are Figure 5.6 plots the 34 cases of measured pillar stress
dependent on the particular coalfield and coal seam. against the ratio w/h collated by Gale (1998) from
Australia, UK and the USA. Gale (1998) recognised
that the strength of a coal pillar is strongly influenced
by its confinement which in turn is influenced by the
strength properties of the strata surrounding the coal.
Hence, in Figure 5.6 three bands are drawn to show
the expected relationship between Cp and w/h for
strong strata, weak strata and for coal seams with clay
filled bedding planes / contacts.
EQUATION 5.13 The relationships shown in in Figure 5.6 were observed
by Seedsman et al. (2005) to bound the empirical
The Constant for the linear form is 2.7 or 3.2 MPa methods, such as those of Equation 5.11, which suggests
for ‘weak coal’ and varies from 4.65 to 4.93 MPa that recognition of confinement offers a means to predict
(average 4.7 MPa) for ‘normal coal’ Salamon et al. pillar strength beyond ratios of w/h > 5.
(2006). Similarly, the r term for the linear form is
0.5176 or 0.6106 for ‘weak coal’ and varies from
0.2591 to 0.5396 (average 0.3670) for ‘normal coal’.
Interestingly, recent work by Mathey and van der
Merwe (2015) found no significant differences in the
strength of laboratory samples of coal from the various
coal seams in South Africa.
Jawed et al. (2013) discusses a dozen equations for
assessing Cp which were published between 1991 and
1992, each of which relates Cp as a function of the ratio
w/h, where the Constant is a parameter for the strength
of a coal specimen with side dimensions of between
30 cm to 1m. They found that the empirical power
relationship proposed by Salamon and Munro (1967),
Equation 5.3, best fits their data for slender pillars from
Indian coal fields. They also repeated the conclusion
from a series of numerical models by Gale (1998) that
the degree of confinement provided to the coal seam is a FIGURE 5.6 Measured pillar stress from various pillar
major factor in determining pillar strength. geometries as presented by Gale (1998)

50
5 | Rock Mass Strength Criteria

Three broad categories of pillar behaviour and failure This comparative observation suggests that the
mode are identified, each defined by an approximate empirical equations for estimating coal pillar strength,
range of width-to-height ratios Mark (1999); at least for failure initiation, implicitly:
Mark (2006). • Include the intact strength of the coal (σc) in their
• Slender pillars, whose w/h < 4, which are subject to constant terms
sudden collapse. • apture the coal pillar’s confinement (σ3) within
C

• Squat pillars, w/h > 10. These pillars can carry very the geometric w/h relationship.
large loads, and may even be strain-hardening. The This is largely in keeping with what Bieniawski
pillar design may still fail because excessive stress is (1968) found that:
applied to surrounding strata.
• Intermediate pillars. These pillars deform until “the stress distribution in pillars, by the nature of
flexure of the overburden transfers some weight their geometry, is very much affected by the width/
away from them. The result is typically a non- height ratio which by producing different stress
violent pillar “squeeze”. concentrations for different pillar geometries, affects
5.3.2.3 Discussion directly the strength of pillars. Consequently, pillar
The current approach to estimate coal pillar strength strength must depend on its width/height ratio”.
is an empirical method based on slender coal pillar
geometry (w/h ≤ 5 or possibly 8) and a characteristic
“strength value” of the particular coal seam. 5.4 PLANE OF WEAKNESS
There are several criteria that explicitly recognise rock
There is a question however, as to the applicability of
mass strength is governed partially by the strength
these empirical methods to wide pillars, i.e. (w/h > 5)
of the intact rock and partially by the strength of
as they are developed from relatively shallow, extensive
the defects within the rock mass. For example, Call
bord and pillar operations for which the pillar was
et al. (2000) provide the following relationships for
designed to hold the weight of overburden.
estimating the Mohr-Coulomb parameters of the
The appreciation that coal under low confinement, rock mass, cm and φm, based on RQD and intact rock
such as in slender pillars or at the edges of pillars in strength.
general, displays brittle behaviour suggests that a DISL
type approach may offer some insight. Equating Cp
with σ1 and then comparing the formulae in Equation
5.4 to Equation 5.13 with that in Equation 5.1 it
follows that:
• The constant term in the equations is an approximate where the subscript m denotes rock mass; i intact rock;
function of intact rock strength. That is (0.3 to 0.5) and j defect parameters and
σc is equivalent to 1.4 to 4.0 MPa, viz:
„„ 6.2 x 0.64 = 4.0 MPa (USA)
„„ 5.12 x 0.56 = 2.9 MPa (Australia)
„„ 2.7 x 0.52 = 1.4 MPa (South Africa – ‘weak coal’)
„„ 
4.8 x 0.37 = 1.8 MPa (South Africa – ‘normal
coal’)
„„ 10.25 x 0.2664 = 2.7 MPa (India)
The equation suffers from the seemingly arbitrary
changes at UCS = 15 and 60 MPa and RQD = 50%
• The geometric term in the equations is an approximate and the inherently scale-dependent RQD.
function of confining stress. That is (1.0 to 2.6)σ3 is
A semi-theoretical approach is provided by
equivalent to 1.19 w/h to 3.35 w/h, viz:
Halakatevakis and Sofianos (2010), (2010) which
„„ 6.2 x 0.54 = 3.35 (USA) extends the plane of weakness theory of Jaeger and
Cook (1979) (Figure 5.7) to consider the strength
„„ 5.12 x 0.44 = 2.25 (Australia) of rock masses crossed by multiple sets of defects of
varying persistence. They proportion the strength of
„„ 2.7 x 0.44 = 1.19 (South Africa – ‘weak coal’)
intact rock (ci, φi) and the strength of the defect (in
„„ 4 .8 x 0.63 = 3.02 (South Africa – ‘normal this method, the instantaneous cd, φd derived from the
coal’) Barton-Bandis strength model for defects) according
to the persistence of the defects. If the persistence is
„„ 10.25 x 0.1514 = 1.55 (India) small, the rock mass strength approaches the intact
rock strength. Conversely if the persistence is large,
the rock mass strength approaches the defect strength.

51
ESTIMATING ROCK MASS PROPERTIES

Apart from estimating the Mohr-Coulomb


parameters for intact rock and for the ‘critical’ defect,
the major difficulty with these methods is the very real
problem of estimating the persistence of the defects
within the rock mass.

FIGURE 5.8 Shear strength relationships for intact rock


through to jointed rock, rock fill and infilled defects as
proposed by Barton, e.g. (Barton 1999; 2014)

FIGURE 5.7 Variation of peak strength as a function of The Barton criterion has the same limitations for
defect dip as defined by Jaeger and Cook (1979); from Brady predicting rock mass strength as it does for intact
and Brown (2005). rock strength. Namely that while it can be fitted to
experimental data to find the shear strength of a rock
mass given the principal stresses at failure, it does not
5.5 BARTON SHEAR find a unique value of σ1 for a given σ3 as its ‘friction
STRENGTH angle’ (ψ) is itself a variable function of σ1 and σ3.
The Barton shear strength criterion, e.g. Barton (1976), Multiple solutions are therefore possible.
Barton and Bandis (1980), Barton (1999) and Barton
(2014) has previously been discussed in relation to
intact rock strength in Chapter 2 and in relation to 5.6 HOEK & BROWN SHEAR
defect shear strength in Chapter 3. The criterion was STRENGTH
empirically derived to originally fit test data on jointed
rock samples and then extended to intact rock and 5.6.1 Criterion
rock masses. It is given as: Hoek et al. (2002) and Brown (2008) state the Hoek-
Brown failure criterion for rock mass as:

where X represents roughness; Y represents the As stated in Chapter 2, for intact rock mb = mi, s
effective compressive strength and Z the basic or =1.0 and a = 0.5, although others note that a should
residual friction (see Figure 5.8). be variable, e.g. (Carter et al. 2007; Carter et al. 2008;
For rock mass strength, the criterion is one or the Diederichs et al. 2007; Mostyn and Douglas 2000;
other of jointed rock or fractured rock. Implicit is You 2011).
the most-likely accurate assumption that rock mass Hoek (1994) stated that the equation is of no
strength is usually governed by defect shear strength. practical value unless the values of the constants m and
Barton (1976), as did Mogi (1966) before him, s can be estimated. In their original work, Hoek and
discussed the concept that the mechanism of rock Brown (1980); (1980), used very limited test work from
mass failure changes as confining stresses increase. At Panguna andesite to derive the constants. The modest
low confining stresses, rock mass failure is brittle and database is reproduced in Table 5.2. Note that the
governed by defect shear strength, as the confinement ‘rock mass’ testing comprised re-compacted 6” (150
increases, rock mass failure becomes ductile. At still mm) and 22.5” (572 mm) diameter samples – that is,
higher confinement, the rock mass strength reaches a still small scale.
limiting value. Singh et al. (2011) working with 154 While relationships between RMR and the
data sets from Sheorey (1997) suggest the confinement parameters m and s were presented in Hoek and Brown
at which this limiting value occurs is approximately (1980); (1980), and are reproduced in Figure 5.9, the
equal to the intact rock’s UCS. equations appear to have been first published in Priest

52
5 | Rock Mass Strength Criteria

TABLE 5.2 Original ‘rock mass’ database from Hoek and Brown (1980); (1980)
Description RMR Number Specimen m mb/mi s
of data diameter
points (inches)
Intact fresh 100 5 2.0 & 4.0 18.9 1.0000 1.00
Undisturbed closely jointed core samples 46 7 6.0 0.278 0.0147 0.0002
Re-compacted samples from a mine bench 28 72 6.0 0.116 0.0061 0
Re-compacted, fresh to slightly weathered, 26 15 22.5 0.040 0.0021 0
rock
Re-compacted, moderately weathered, rock 18 5 22.5 0.030 0.0016 0
Re-compacted, highly weathered, rock 8 3 22.5 0.012 0.0006 0

and Brown (1983) as:

The equations were re-assessed to be applicable to


‘ disturbed rock masses, such as those loosened by poor
blasting practice, and those in embankments or waste
dumps’ and simplified in a subsequent publication by
Brown and Hoek (1988), to:

FIGURE 5.9 Original plot showing the data on which


That same publication also introduced the following the relationships between RMR and the parameters mb
equations for peak strengths of undisturbed rock & s are based, from Hoek and Brown (1980); (1980). The
relationship between RMR and s, were based on 2 points –
masses, though no details of the cases that this was
s=0.002 and s=1. Superimposed on the original plot are the
based on were provided: equations proffered by Priest and Brown (1983) and Brown
and Hoek (1988). Blue lines are for mb /mi relationship; red
lines are for the s relationship.

Hoek (2007) found the correlation between GSI and


All three sets of equations are superimposed onto the RMR to be problematic for poor quality rock masses
original plot reproduced in Figure 5.9. The differences and recommended that the GSI be estimated directly by
between the ‘undisturbed equations’ of Brown and means of its charts, which are described in Chapter 4.
Hoek (1988) and the original plot, the equations of Following the work of Sonmez and Ulusay (1999)
Priest and Brown (1983), the ‘disturbed equations’ of who suggested including a factor to explicitly account
Brown and Hoek (1988), can clearly be seen. for the disturbance caused by the excavation method,
RMR was replaced with GSI in Hoek (1994), who and other case studies, Hoek et al. (2002) introduced
also separated undisturbed rock masses into two a Disturbance Factor, D. They gave general guidelines
groups at the value of GSI = 25. to assess D between 0 for undisturbed rock masses to
1 for very disturbed rock masses. This enabled them
to combine the ‘undisturbed’ and ‘disturbed’ sets of
equations. They also removed the step at GSI = 25
and introduced the relationship between GSI and the
parameter a to give the relationships in Equation 5.14.

Here the GSI is equal to the 1976 version of RMR


(Hoek 1994). Hence, these equations for mb /mi and s
are the same as those given above by Brown and Hoek
(1988). Hoek (1994) also allows GSI to be calculated
using the 1989 version of RMR as GSI = RMR89 – 5. EQUATION 5.14

53
ESTIMATING ROCK MASS PROPERTIES

From the above presentation, it is clear these In keeping with those limitations, Marinos and
relationships in Equation 5.14 are a development Hoek (2000) state that GSI “should not be applied
from the original work by Hoek and Brown (1980). to those rock masses in which there is a clearly defined
It can be readily seen that as D varies from 0 to 1 the dominant structural orientation”. Many sedimentary
equations vary between the ‘undisturbed rock mass’ rock masses could fit this description. However,
and ‘disturbed rock mass’ cases of Brown and Hoek Marinos and Hoek (2000) in discussing anisotropic
(1988). rock masses go on to state that “ it is sometimes
In relation to D, Hoek et al. (2002) noted that appropriate to apply the Hoek-Brown criterion to the
“a large number of factors can influence the degree overall rock mass and to superimpose the discontinuity as a
of disturbance in the rock mass surrounding an significantly weaker element.” Further, they specifically
excavation and that it may never be possible to quantify published GSI charts for flysch, molasse, limestone,
these factors precisely”. Hoek and Diederichs (2006) ophiolites and gneiss. The implication is that rock
elaborated further by stating that D “will vary for each mass strength parameters for anisotropic rock masses
application, depending upon the excavation and loading can be established based on the GSI approach provided
sequence for the particular structure being designed.” defects are explicitly modelled.
It is also worth noting that because of this history Diederichs (2007) suggests that the GSI / Hoek-
the relationships in Equation 5.14 have their basis as Brown approach is of limited reliability when used for
the very limited test work from Panguna andesite ‘rock spall-prone rock masses which he defines as mi > 15
masses’, shown in Figure 5.9. and GSI > 75 and may be of mixed success when GSI
is equal to 65-75.
5.6.2 Limitations Brown (2008) opines that there is a limit to the
Several limitations of the Hoek-Brown failure criterion range of GSI values over which the Hoek-Brown
were identified during the criterion’s development, criterion may be applied: ‘special care must be exercised
e.g. Hoek (1994); Hoek and Brown (1980); Hoek and at low values of GSI of below about 30’ or ‘ σc < 15 MPa’.
Brown (1997). The UCS limit of 15 MPa is also agreed with by others
(Carter et al. 2007; Carter et al. 2008; Carvalho et
• The criterion is only applicable to intact rock or to al. 2007; Dinc et al. 2011). This is interesting given
heavily jointed rock masses which can be considered
homogenous and isotropic, as summarised in Figure that one of the primary reasons for Hoek (1994)
5.10. It should not be used in cases in which the introducing the GSI was the inability of RMR to deal
strength of the rock or the rock mass is likely to be with poor quality rock masses.
dominated by one or two sets of discontinuities, or Further, Carter et al. (2007); (2008); Carvalho et
by a major through-going structure such as a fault al. (2007) state difficulties are experienced with the
or shear zone. Hoek-Brown criterion at the two ends of the rock
• In highly schistose rocks the criterion applies to the competency scale largely because the behaviour of
intact rock components only; the strength of the these rock masses becomes less controlled by defects.
discontinuities need to be explicitly considered. Carter et al. (2007) summarise their research, which
is also published elsewhere (Carvalho et al. 2007;
Diederichs 2007), and suggest that the Hoek-Brown
material parameters, mb, s and a for very low strength
rocks and spall-prone rocks are functions of σc :
• For very low strength rocks (UCS < 10 MPa) where
rock mass behaviour is matrix controlled, the rock
mass can be considered to be homogeneous soil-like

FIGURE 5.10 The influence of scale on the type of rock


mass behaviour model (Hoek and Brown 1997)

54
5 | Rock Mass Strength Criteria

• For spall-prone rock masses (mi > 15 and GSI >>


65), where rock mass failure demands creation of
new fractures; essentially unjointed, hard, brittle
rock masses

FIGURE 5.11 Variation of Hoek-Brown parameters for


where UCS* is where stress-induced fracturing spall-prone rocks with mi = 20 (Diederichs et al. 2007)
initiates, i.e. 0.3 to 0.5 σc . The residual values are
included, as the peak strengths by themselves result
While these developments are moving away from
in parameters that ‘will significantly over-predict
Hoek & Brown’s stated objective of a simple criterion,
failure if used in non-linear plasticity codes’ (Diederichs
et al. 2007). they highlight concerns with the criterion’s application.
The variation of the Hoek-Brown peak parameters for If one accepts the observations of Diederichs (2007),
spall-prone rocks with respect to GSI as recommended Carter et al. (2007), (2008) Carvalho et al. (2007) and
by Diederichs et al. (2007) is shown in Figure 5.11 for Brown (2008) then the range over which the ‘standard
the particular case of mi = 20. GSI approach’ should be used to predict the Hoek-

FIGURE 5.12 The range suggested by researchers mentioned in the text, over which GSI can be used to predict the Hoek-
Brown strength parameters of a rock mass.

55
ESTIMATING ROCK MASS PROPERTIES

Brown strength parameters is limited to approximately the frictional strength progressively mobilises and
30 – 65 (Figure 5.12). replaces the cohesive strength component. Martin and
Perhaps the major issue with the GSI / Hoek-Brown Chandler (1994), You (2011) and Barton and Pandey
criterion approach, and perversely perhaps its major (2011) also argue the independence of cohesion and
attraction, stems from the user-friendliness of the once frictional strength.
freeware program RocLab. A strength envelope can be While Hajiabdolmajid et al. (2002) were able to
generated in RocLab knowing nothing about the rock match the shape of the failure zone of the Mine-By
mass except a loose geological description. Experiment with the CWFS model, the model predicts
The following guidelines should be considered an ‘upward-bent failure envelope’ in shear versus normal
for numerical modelling rock masses defined by stress plot rather than the typical downward-bent or
GSI (Brown 2008) assuming that the Hoek-Brown linear failure envelope (see Figure 5.13). It may be
criterion is applicable. that the failure envelope for intact rock is S-shaped as
• Rock masses with GSI < 30 should be modelled as suggested by Diederichs (2007), see Figure 5.2.
perfectly plastic materials with zero dilation angle
• Rock masses with GSI > 65 should be modelled 5.8 SRM
as elastic, brittle materials with a dilation angle
approximately equal to φ/4 Brown (2008) suggests the development of “practically-
• Other rock masses should be modelled as useful numerical methods” may provide a means of
strain softening material with a dilation angle estimating the mechanical properties of rock masses.
approximately equal to φ/8. Numerical modelling of Synthetic Rock Mass
(SRM) is being researched to quantify rock mass
behaviour at large scales in order to address the
5.7 CWFS limitations of the classification systems used to factor
Hajiabdolmajid et al. (2002) found that continuum intact rock strength into rock mass strength, e.g. Mas
models with traditional failure criteria (e.g. Hoek– Ivars (2010). In addition to those described previously,
Brown or Mohr-Coulomb) which implicitly assume the the limitations of classification systems include
simultaneous mobilisation of cohesive and frictional
the limited ability to consider strength anisotropy
strength components were not successful in predicting
(resulting from a preferred joint fabric orientation) and
the extent and depth of brittle failure in massive Lac
strain softening/weakening.
du Bonnet granite at the Atomic Energy of Canada’s
(AECL’s) Mine-By Experiment at the Underground An SRM is a numerical technique to represent a
Research Laboratory (URL) in Manitoba, Canada. jointed rock mass. It does this by combining an assembly
They also note that Schmermann and Osterberg (1960) of particles bonded together at their contacts with an
showed that even in cohesive soils, c and φ might not embedded discrete fracture network (DFN). Assembly
necessarily be mobilised simultaneously. of particles can be modelled as bonded particles, e.g.
They adopted a continuum model with strain- Cundall et al. (2008) or as discrete elements as in Lan
dependent cohesion weakening – frictional et al. (2010) who modelled individual mineral grains
strengthening (CWFS) criterion. The characteristics of (Figure 5.14).
which are that at the early stage of failure, the mobilised The behaviour of an SRM depends on both the creation
strength is mainly due to the cohesive strength of new fractures through intact material and the slip/
component and as inelastic strains accumulate, opening of pre-existing joints.

FIGURE 5.13 Cohesion-weakening and frictional-strengthening (CWFS) as functions of plastic strain for Lac du Bonnet
granite (Hajiabdolmajid et al. 2002)

56
5 | Rock Mass Strength Criteria

correlations; the major impetus for developing the


SRM method.

5.8.2 Limitations
Hoek and Martin (2014) list the followng requirements
of a numerical rock mass model as noted by Potyondy
and Cundall (2004).
• Continuously nonlinear stress-strain response, with
ultimate yield, followed by softening or hardening
• Behaviour that changes in character, according to
stress state
• Memory of previous stress or strain excursions, in
both magnitude and direction
• Dilatancy that depends on history, mean stress and
initial state
FIGURE 5.14 Grain-sized UDEC model for UCS tests on Lac
• Hysteresis at all levels of cyclic loading/unloading
du Bonnet granite and Aspo Diorite; from Lan et al. (2010) • Transition from brittle to ductile shear response as
the mean stress is increased
5.8.1 Calibration • Dependence of incremental stiffness on mean stress
The properties obtained from standard UCS tests on and history
core samples are scaled to a size that will be modelled • Induced anisotropy of stiffness and strength with
in the SRM as an intact rock block (the SRM block). stress and strain path
The scaling makes use of the empirical relations • Nonlinear envelope of strength
proposed by Hoek and Brown (1980) or by Yoshinaka • Spontaneous appearance of microcracks and
et al. (2008). The properties of the SRM block are then localized macro fractures
calibrated to these scaled values (UCS, average Young’s • Spontaneous emission of acoustic energy.
modulus and Poisson’s ratio). Limitations of the SRM technique as identified by
Yoshinaka et al. (2008) however, concluded that Mas Ivars (2010) include the following.
the strength reduction is dependent on rock type, • The method does not reproduce the σc / σct ratios
strength, texture, etc., i.e. it is not constant (Figure observed in hard rocks
5.15). Ongoing research at the University of New • e intersection between joint planes in the DFN
Th

South Wales, e.g. Masoumi et al. (2012) is confirming follows a hierarchical model, which significantly
the work by Hawkins (1998) that suggests for some affects the rock mass properties of the SRM
rocks, notably sedimentary rocks, UCS decreases, not • The predictions of rock mass behaviour obtained
increases, as core size becomes smaller than 50 mm using the SRM approach are only as good as the
diameter (Figure 5.16). This implies that part of the representation of the insitu joint network by
current SRM approach relies on imprecise empirical the DFN

FIGURE 5.15 Scaling of UCS values to block size (Mas Ivars FIGURE 5.16 UCS tests on seven different sedimentary
2010) rocks conducted by Hawkins (1998)

57
ESTIMATING ROCK MASS PROPERTIES

Further, a major disadvantage is the extensive Despite its limitations, the Hoek-Brown shear
experimental validation and calibration required to verify strength criterion remains the most widely used to
that models are capturing the observed macroscopic assess rock mass failure. On balance of published
behavior of rock (Lisjak and Grasselli 2014). material, it may predict the overall size of failure
but it is not particularly accurate in replicating
observed failure.
5.9 CONCLUSIONS Further, the Hoek-Brown shear strength criterion
The ideal rock mass failure criterion would satisfy the for rock masses should be assessed using the GSI
following main requirements, paraphrasing Hoek and methodology, noting it is better suited for the range
Brown (1980): GSI = 30 – 65.
• Be simple criterion developed on a sound theoretical Based on the discussions presented in this chapter,
basis
the following criteria are assessed to provide the best
• Describe the response of an intact rock sample from ways for designers to estimate rock mass strength.
uniaxial tensile stress to true triaxial, compressive These criteria are looked at in subsequent chapters.
stress in ductile or brittle failure
• Brittle rock failure (or ‘damage initiation’)
• Predict the influence of one or more defect sets • Hoek-Brown shear strength.
upon the rock behaviour
• Project the behaviour of a full-scale rock mass
containing several sets of defects

58
6 | Rock Mass Modulus

6 ROCK MASS
MODULUS

6.1 OUTLINE • 
Measurements of tunnel convergence engage
There appear to be five broad approaches to estimating 1000 m or more of rock mass in varying states of
3

rock modulus (5). stress and strain.


1 Correlations with index tests such as those • The dynamic modulus (small strain, small load)
presented by Tziallas et al. (2009). Chief of these, derived from seismic velocity needs to be calibrated
is the correlation of intact rock modulus, Ei , with to a static modulus.
intact unconfined compressive strength, σc , via the
modulus ratio, MR = Ei /σc . The MR, which was
introduced by Deere (1968), is considered to vary
with rock type and texture (Figure 6.1). The original
range of MR was updated by Hoek and Diederichs
(2006) to extend from 150 to 1000, with typical
values of 300 to 500.
2 Correlations with rock mass classification systems.
3 Analytical or numerical techniques which require
values for Ei , and for joint stiffness (kn, ks) to estimate
the equivalent modulus of the rock mass, Em.
4 Analytical equations that calculate modulus for an
equivalent continuum.
5 Back-analysis using displacement data measured
during excavation.
The state-of-practice in predicting the modulus
of rock masses relies on correlations developed from FIGURE 6.1 Graph showing the variation of intact Young’s
Modulus to intact Uniaxial or Unconfined Compressive
field case studies where the in situ modulus has been Strength, after Deere (1968). Accessed on 7 October
measured (Hoek and Diederichs 2006). 2014 http://lmrwww.epfl.ch/en/ensei/Rock_Mechanics/
In situ testing for rock masses includes plate- ENS_080312_EN_JZ_Notes_Chapter_4.pdf
bearing testing, flat-jack tests, borehole dilatometer,
back-analysis of tunnel convergence, seismic velocity, The scale effect of reducing modulus with increasing
etc. However, each test measures a slightly different rock volume is well documented, for example Galera
modulus, that being for a different volume of rock, et al. (2005). The ideal is to select appropriate design
in different directions and at different strains. values of modulus that correspond to the expected
For example: range of deformation for the expected volume of rock
• A borehole dilatometer or pressuremeter engages mass. Figure 6.2 schematically shows the variation of
approximately 0.1 m3 of rock. shear modulus, G [=E/2(1+v)] with respect to shear
• Because most boreholes tend to be vertical or near strain and various in situ tests.
vertical, pressuremeter tests typically measure This chapter presents published in situ rock mass
horizontal modulus, which in sedimentary rocks modulus data before describing and appraising several
may be higher than vertical modulus. classification systems, analytical and numerical
• The vertical modulus is typically measured in techniques to estimate rock mass modulus. The back-
laboratory testing of borehole core samples (i.e. analysis method for estimating in situ rock mass
UCS testing). modulus is developed in Chapter 7.

The terms Young’s modulus and Poisson’s ratio apply to linear elastic materials. In rock, they are often given as a tangent modulus at a stress level close to
(5) 

50% of the UCS or as a secant modulus up to the UCS. However, the term rock modulus is also used for the deformation modulus of a rock mass.

59
ESTIMATING ROCK MASS PROPERTIES

FIGURE 6.2 Variation of modulus with shear strain (Geotechnical Engineering Circular #5, 2002). Though this curve was
prepared for use with soils, the same concept applies to rock masses.

6.2 MODULUS RATIO The results comprise 262 tests of Hawkesbury


The UCS with modulus laboratory test results which Sandstone, 34 of Gosford Sandstone, 29 of Ashfield
were undertaken for six recent tunnelling projects in Shale and 4 of dolerite. The results are plotted on
Sydney – the Epping to Chatswood Rail Link, Lane Figure 6.3 and Figure 6.4 and show that most fall with
Cove Tunnel, the CBD Metro (which was not built), the medium modulus ratio band as defined in Figure
the Northwest Rail Link, the M4East and the Sydney 6.1, i.e. a MR between 200 and 500, although there are
Metro is used. In addition, the testing carried out by some results that are have a low ratio and are between
Masoumi (2013) on Gosford Sandstone is also used. approximately 100 and 200.

100

PRL
PRL - oven dry
LCT
E [GPa]

10 CBD Metro
NWRL
Sydney Metro
M4E
Gosford Sandstone
Dolerite

1
10 100
UCS [MPa]

FIGURE 6.3 Intact modulus plotted against UCS. Open symbols (most of the data) represent sandstone; solid symbols
represent shale. There are four data points of dolerite, sampled from dykes.

60
6 | Rock Mass Modulus

FIGURE 6.4 Intact modulus plotted against UCS for Ashfield Shale (top) and Hawkesbury Sandstone (bottom). These graphs
show the same data as Figure 6.3.

61
ESTIMATING ROCK MASS PROPERTIES

This is in keeping with the Hawkesbury Sandstone 6.3 DATA


data presented by Robson (1978) in Figure 6.5 and
with the Ashfield Shale data presented by Won (1985) 6.3.1 Reliability
in Figure 6.6. It is important to understand some of the difficulties
of in situ testing before reviewing published rock mass
moduli data.
1 The test methodology strongly influences the test
result.
„„ 
Bieniawski (1978) noted that the modulus as
interpreted from the best understood test –
the plate-bearing test, which simply involves
applying a load to the rock surface and measuring
the deformations – can still differ by a factor of
two to three.
2 Bieniawski (1978) then went on to make several
recommendations for in situ investigations; the
foremost two of which are paraphrased here.
„„ 
The first is that the in situ rock mass modulus
should be estimated by a sufficient number
of at least two different direct types of tests.
Bieniawski preferred the plate-bearing and
Goodman borehole jack tests.
What is meant by ‘a sufficient number’ depends
on the geological assessment of the rock mass. A
straightforward rock mass will have fewer tests
than a complex rock mass.
„„ 
The second recommendation is that the in situ
stress field needs to be estimated, as it was found
to affect plate-bearing test results.
Though Bieniawski recommended using the
FIGURE 6.5 Intact modulus plotted against UCS data for Goodman borehole jack, 28 years later Hoek and
Hawkesbury Sandstone from Robson (1978) as reproduced Diederichs (2006) found that the least reliable in situ
in Pells (2004). measurements are those from various down-hole jacks
and borehole pressure-meters. They concluded that the
measurements obtained from such devices are difficult
to interpret, particularly in hard, jointed rock masses
in which the stressed rock volume is obviously small.
It is also supported by the data presented in Chun et
al. (2009), which comprises 61 points where the mass
modulus of very poor quality (RQD ≤ 20%) but high
strength rock (UCS ranged between 12 to 250 MPa)
was obtained from borehole pressuremeter tests (see
Figure 6.7).
The difficulty interpreting in situ rock mass modulus
testing was also expressed by Vibert and Ianos (2015)
who assessed the results from 199 borehole dilatometer,
33 flat jacks, 149 plate loads and seismic refraction tests
(called the Scarabee method) at six large dam projects
in France in a variety of rock types. They concluded: ‘As
a common feature of these tests [the borehole dilatometer,
flat jacks and plate load tests], it can be said that they
are relatively difficult to implement, and interpretation
and calculation of moduli necessitate assumptions which
cannot be directly verified. Due to the natural dispersion
FIGURE 6.6 The intact modulus plotted against UCS data of results, they are obviously to be performed in large
for Ashfield Shale from Won (1985). number for allowing assessing an average behaviour.’

62
6 | Rock Mass Modulus

An additional complexity is that many of the However, the inherent scatter in measured Em as
methods to predict rock mass modulus assume discussed in the preceding Section 6.3, means that any
isotropic conditions. In sedimentary rocks and schists, curve fitting approach is expected to be a poor predictor.
the modulus is evidently dependent on the direction As discussed in Chapter 4, the most widely used
of loading. For example, Tziallas et al. (2009) quote classification systems for underground excavation
an MR range of 250 to 1100 for schist, dependent on design are Q (Barton et al. 1974; Grimstad and Barton
relative orientation of testing to foliation, compared 1993) and GSI (Hoek 1994; Hoek and Brown 1997;
to the typical range from most rocks of between 300 Marinos and Hoek 2000) and by antecedence RMR
and 500. (Bieniawski 1973; Bieniawski 1976; Bieniawski 1989).
The reasonable conclusion from this discussion is The correlations between E m and these systems and
that the in situ test methodology – multiple direct tests, between E m & RQD (Deere 1968) are discussed in the
the influence of test methods, in situ stress – needs to following sections.
be considered when assessing the reliability of in situ
rock mass modulus. If the methodology used to report 6.4.1 RQD
in situ data in the literature is not known or explicitly Zhang and Einstein (2004) suggested the following
included in the interpretation, then how reliable is the correlations between RQD and modulus, which they
modulus? More fundamentally, this also questions claimed provide useful bounding values (Figure 6.7).
the logic in choosing one predictive technique over
another based on statistical correlations of published • Lower bound E m ⁄ Ei=0.2×100.0186RQD-1.91
data of unknown reliability. • Mean E m ⁄ Ei=1.0×100.0186RQD-1.91
A further conclusion is that the more reliable • Upper bound E m ⁄ Ei=1.8×100.0186RQD-1.91
estimates of in situ rock modulus are those derived from
tests that engage larger volumes of rock; stepping from
flat jack and plate load testing to seismic refraction Figure 6.7 however shows the large scatter in the ratio
tests and back-analysis of convergence data. E m ⁄ Ei compared to RQD and the three correlations.
The figure suggests that the correlations are probably
6.3.2 Sources not sufficiently accurate for meaningful engineering
There are five main publications that report in situ design. The ellipse in Figure 6.7 highlights data that
rock mass modulus data – Bieniawski (1978); Hoek shows the unreliability of borehole pressuremeter tests
and Diederichs (2006); Palmström and Singh (2001); obtained in very poor quality (RQD ≤ 20%) but high
Serafim and Pereira (1983); Stephens and Banks strength rock (UCS ranged between 12 to 250 MPa)
(1989). Of course, there are other publications which (Chun et al. 2009).
present particular site data, such as Ajalloeian and
Mohammadi (2014); Chun et al. (2009); Kallu et al.
(2015); Kayabasi et al. (2003); Nejati et al. (2014);
Vibert and Ianos (2015) to name a few, but most of the
classification-based correlations appear to use the five
main publications listed above.

6.4 CLASSIFICATION
SYSTEMS
According to Palmström and Singh (2001) reasonable
estimates of E m can be made using classification
systems for moduli in the following ranges, where E m
is in GPa.
• R MR For 55 < RMR < 90
E m = 2RMR 100 (Bieniawski 1978)
 For 30 < RMR < 55
E m = 10(RMR -10)/40 (Serafim and Pereira
1983)
• Q For 1 < Q0.4< 30
E m = 8Q
• In
addition, for massive or slightly jointed rock FIGURE 6.7 Correlation between RQD and Em/Er, after Zhang
mass, E m [GPa] ≈ 0.2σc [MPa]. (2010); Zhang and Einstein (2004).

63
ESTIMATING ROCK MASS PROPERTIES

TABLE 6.1 The tests carried out to assess the rock mass modulus reported in Bieniawski (1978)
TEST ORANGE RIVER DRAKENSBERG ELANDSBERG

Small flat jack - - 19

Large flat jack - - 3

Goodman borehole jack - - 45

Plate-bearing tests 53 14 12

Pressure chamber / 5 27
Tunnel relaxation measurements

Geophysics profiles 40 32 68

In situ stress measurements 38 30 23

6.4.2 RMR Though Bieniawski (1978) does not explicitly state it,
the data presented in Figure 6.8 appears to come from
6.4.2.1 Bieniawski
the more reliable in situ test methods, viz: Elandsberg
Bieniawski (1978) presented the rock mass modulus
– plate bearing, flat jacks, tunnel relaxation, seismic;
data derived from plate-bearing tests, flat jack tests,
Drakensberg – plate bearing, seismic; Orange River
borehole dilatometer tests and geophysical tests at
– plate bearing, pressure chamber, tunnel relaxation;
three engineering projects in South Africa: the Orange
Witbank – in situ coal pillar compression tests; Le
River Water Project, the Drakensberg Pumped Storage
Roux Dam – plate bearing; Dinorwic – flat jacks; and
Scheme and the Elandsberg Pumped Storage Scheme.
Gordon Scheme – flat jacks. Bieniawski (1978) does
The extensive in situ testing is summarised in Table
state that Equation 6.1, which is shown in the graph in
6.1. The important rock mass features of the three sites
Figure 6.8, has an accuracy to predict in situ rock mass
taken from Bieniawski (1978) are summarised in Table
modulus of better than 20%.
6.2. Bieniawski (1978) presented the data on a graph
correlating the interpreted rock mass modulus with
the inferred RMR as reproduced in Figure 6.8 which
also has the in situ modulus from the South African
Witbank coal field. EQUATION 6.1

FIGURE 6.8 Measured in situ rock mass modulus of deformation against RMR76 (Bieniawski 1978)

64
6 | Rock Mass Modulus

TABLE 6.2 Case histories reported in Bieniawski (1978)


DESCRIPTION RQD DEFECTS RMR76 Ei [GPa] Em [GPa]
[%]

Dolerite 80 – 100 Excellent condition 75 – 85 63.4 – 77.2 23.5 – 31.8


Orange River Project

Highly Unweathered, horizontal


homogeneous foliation
spaced 1-1.7 m
Some vertical jointing
spaced 8-10 m
Very slight water inflow

Mudstone, ? Closely spaced, horizontally 41 – 70 13.3 ± 6.5 (v)


Siltstone, bedded 17.8 ± 8.3 (h)
Sandstone

Mudstone, 80 – 100 Unweathered, closely spaced 46 – 55 19.4 (v) 18.3 (v)


Drakensberg

Siltstone, near horizontal bedding & 28.5 (h) 25.0 (h)


Sandstone Some minor jointing 62 – 67
No groundwater

Greywacke 75 – 85 Sub-vertical foliation 66 – 80 73.4 ± 3.8 40.1 ± 14.1


Elandsberg

Two joint sets plus minor


faulting

Phyllite 65 – 70 Sub-vertical foliation 43 – 60 56.0 ± 11.9 18.7 ± 6.5


Two joint sets plus minor
faulting

6.4.2.2 Serafim and Pereira Serafim and Pereira (1983) do not discuss the
Serafim and Pereira (1983) brought attention to the test methods used to collect their data from the
evident issue with Equation 6.1 and Figure 6.8 that 15 test sites. There are also no details, e.g. defect
a negative in situ modulus is predicted for RMR76 ≤ characteristics, intact strength, etc., of the rock masses
50. They added data from 15 test sites at seven dams to independently derive RMR76.
to Bieniawski’s graph and proposed the exponential
relationship of Equation 3 to predict in situ modulus
for poor quality rock masses. The relationship is shown
in Figure 6.9. EQUATION 6.2

FIGURE 6.9 Measured in situ rock mass modulus of deformation against RMR76 as published in Serafim and Pereira (1983).
The points shown as ‘+’ were taken from Bieniawski’s database.

65
ESTIMATING ROCK MASS PROPERTIES

6.4.2.3 Stephen and Banks Stephens and Banks (1989) also reported the
The in situ deformation modulus of sandstone and measured modulus from 32 borehole jacking tests(7).
conglomerate as measured from eight in situ plate- They presented the results in the graph shown in
bearing tests at the Portugues dam site in Puerto Rico Figure 6.11. The values labelled as Ecal (calculated
are presented in Stephens and Banks (1989). The plate- modulus from the linear portion of the deformation
bearing tests generally followed the standard ASTM D versus hydraulic pressure curve) have been digitised to
4395-84 (6). The modulus values were presented along produce the 23 data points in Table 6.4. No in situ
with the corresponding estimated RMR76 (Table 6.3). stress measurements were carried out at the dam site.
The results are summarised in Figure 6.10. The two There are also no details, e.g. defect characteristics,
points that lie well below the Serafim and Pereira (1983) intact strength, etc., of the rock mass to independently
RMR relationship shown in the figure are the tests from derive RMR.
the shear zone and the damaged rock surface.

TABLE 6.3 In situ results from plate bearing tests reported in Stephens and Banks (1989)
DESCRIPTION RMR76 Em [GPa]

Conglomerate (CG) / sandstone (SS) 66 29.6

Conglomerate / sandstone 62 17.9

Sandstone (damaged rock surface) 43 2.8

Sandstone 63 17.2

Sandstone 62 34.5

Sandstone / dyke 38 4.8

Sandstone / dyke 49 8.3

Shear 44 2.1

FIGURE 6.10 Measured in situ rock mass modulus from FIGURE 6.11 Deformation modulus obtained from
plate bearing tests against RMR76 from Stephens and Banks Goodman borehole jacking tests (EGJ) against that derived
(1989). The Serafim and Pereira (1983) relationship in with the Serafim and Pereira (1983) RMR76 relationship
Equation 6.2 is shown. (ERMR) (Stephens and Banks 1989).

(6)
ASTM D 4395-84 recommends using 0.5 to 1 m diameter plates
(7)
The tests used a Goodman jack modified to fit in a 96 mm diameter hole instead of 76 mm

66
6 | Rock Mass Modulus

TABLE 6.4 Goodman borehole jack tests results digitised from Figure 6.11 Stephens and Banks (1989)
EMPIRICAL RELATIONSHIP IN SITU MODULUS
Em Em

[psi x106] [GPa] RMR76 [psi x106] [GPa]

1.4 9.65 49 3.3 22.8

1.77 12.20 53 4.43 30.5

2.05 14.13 56 1.1 7.6

2.15 14.82 57 1.7 11.7

2.55 17.58 60 2.55 17.6

2.55 17.58 60 4.4 30.3

2.62 18.06 60 2.83 19.5

2.65 18.27 60 3.34 23.0

3.2 22.06 64 2.18 15.0

3.2 22.06 64 6.12 42.2

3.24 22.34 64 3.68 25.4

3.7 25.51 66 2.9 20.0

4.3 29.65 69 2.75 19.0

4.3 29.65 69 3 20.7

4.92 33.92 71 2.98 20.5

5.7 39.30 74 2.9 20.0

5.8 39.99 74 2.3 15.9

5.81 40.06 74 4.1 28.3

5.82 40.13 74 3.9 26.9

5.9 40.68 74 1.4 9.7

9.18 63.29 82 4.12 28.4

9.2 63.43 82 5.22 36.0

9.35 64.47 82 1 6.9

6.4.2.4 Palmström and Singh data obtained from the drill & blast adit by three
A comparison between the in situ modulus derived to plot it in Figure 6.12. Given the 2001 date of the
from plate bearing tests carried out within a carefully publication, it is assumed that the modulus data is
excavated adit and within a drill & blast adit was made plotted against RMR89.
by Palmström and Singh (2001). They concluded that
6.4.2.5 Other
the modulus of the drill & blast affected rock mass
could typically be a third of that obtained from the Various correlations with RMR were derived and
undisturbed rock mass in the carefully excavated adit. modified by subsequent researchers. For example,
Figure 6.13 presents the approximate relationship
Palmström and Singh (2001) presented the data,
between Em and the ratio Em /Ei with RMR 76 as given by
which comprised 42 test sites at eight hydropower
Galera et al. (2005). The database is largely the same as
projects in India, Nepal and Bhutan in gneiss,
granite, mica schist, sandstone, mudstone, siltstone, that used for the correlation with RQD in Figure 6.7.
and dolerite/hornblende rhyolite, in Figure 6.12. The Mohammadi and Rahmannejad (2010) compared
UCS varied from 30 to 230 MPa. To address the issue five separate, published correlations between RMR and
of blast damage reducing the rock mass modulus, rock mass modulus with a dataset of 171 case studies
Palmström and Singh (2001) multiplied the modulus and concluded that the best estimate was in fact a

67
ESTIMATING ROCK MASS PROPERTIES

FIGURE 6.12 Deformation modulus obtained from plate bearing tests from Palmström and Singh (2001). The data
comprises undisturbed rock mass (2 points plus that from Stripa mine) plus disturbed rock mass (40 points from drill & blast
adits). It is assumed that RMR here is RMR89.

cubic equation Em = 0.0003RMR 3 – 0.0193RMR 2 +


0.3157RMR + 3.4064. Even this arbitrary and rather
awkward relationship was reported to have only a
modest correlation of r2=0.84. Further, there is no data
in Mohammadi and Rahmannejad (2010) to indicate
whether the RMR and Em were reliably measured in
any or all of the 171 case studies.
Nejati et al. (2014) provides in situ modulus as
obtained from eight plate load tests and 44 borehole
dilatometer tests from the mudstones, siltstones,
sandstone and conglomerates of the abutments of the
Gotvand dam site in Iran.
They provide estimates of RMR for each test site
(see Figure 6.14) but unfortunately do not distinguish
between data from the plate load tests and those from
the borehole dilatometer tests. Hence, the reliability of
the data could not be assessed, though it is probable
that the rock mass in the abutments was disturbed due
to blasting and stress relief.

FIGURE 6.13 Rock mass modulus (top) and ratio of rock


mass to intact rock modulus (bottom) as a function of RMR
(Galera et al. 2005)

68
6 | Rock Mass Modulus

Grimstad and Barton (1993) imply the data assigned


to “Q case records“ in Figure 6.15 come from numerical
back-analysis of measured deformations though
they do not state it. There are no details, e.g. defect
characteristics, intact strength, etc., of the various rock
masses to independently derive Q. The degree to which
the rock mass is disturbed is not known.
Barton (1995) updated the correlation as:

EQUATION 6.3

FIGURE 6.14 In situ modulus against RMR for the Gotvand


where M is the rock mass deformation modulus. This
dam site from Nejati et al. (2014) is also shown in Figure 6.15.
The UCS normalised Q-value, Qc, was introduced in
Barton (1995) to improve the correlation of rock mass
quality and seismic velocity. The Qc value is obtained
by normalising to nominal hard rock compression
6.4.3 Q strength value of 100 MPa (Equation 6.4).
A correlation between Rock Mass Quality, Q, and
modulus was provided by Grimstad and Barton (1993)
as Em = xlog10Q. They gave the typical value for x as
25 but noted that it could range between 10 and 40. EQUATION 6.4
These three correlations are shown in Figure 6.15.
Publications subsequent to Barton (1995), such
as Barton (2002) and Barton and Grimstad (2014),
further clarified the use of Qc to improve correlation
between rock mass deformation modulus and seismic
velocity. Hence, the relationship is Equation 6.5:

EQUATION 6.5

A similar relationship was derived by Palmström


and Singh (2001), i.e. E m = 8Q 0.4.

6.4.4 GSI
Hoek and Diederichs (2006) collated a large database
of what they claim to be reliable in situ measurements,
carried out between 1960 and 2005, of rock mass
modulus of sedimentary, igneous and metamorphic
rocks from China (457 data points) and Taiwan (37
data points).
• 18 cases where the in situ modulus was estimated
using back analysis of convergence data
FIGURE 6.15 Correlation between rock mass modulus and • 53 cases where flat jacks were used
Q (Barton 1995) • 423 cases which used plate tests.
There will be some inherent variability in the
measured rock mass modulus values in this eclectic
database, which suggests that the difference between
predicted and actual rock mass modulus is probably
larger than the 10% error claimed by Hoek and
Diederichs (2006) for their correlation.

69
ESTIMATING ROCK MASS PROPERTIES

Figure 6.16 presents the Hoek and Diederichs


(2006) data against their estimated GSI for those
rock masses. The GSI was estimated directly from
RMR, which is problematic for poor quality rock as
found by Hoek (2007) who recommended that the
GSI be estimated directly by means of its charts (see
Chapter 4). It is apparent from Figure 6.16 that rock
mass modulus increases with increasing GSI but so
does its variability as judged by the scatter in the data.
Hoek and Diederichs (2006) attribute the scatter in
the data to “ inherent scatter in the values of GSI, intact
rock properties, rock mass modulus Em, and the effects of
disturbance due to blasting and/or stress relief.”
The last item is underlined to highlight the fact
that the degree to which the Chinese and Taiwanese
data represents disturbed rock mass is not known. The FIGURE 6.16 Measured rock mass modulus of deformation
number of data points in the various GSI ranges was against GSI for the Chinese and Taiwanese data from Hoek
counted as shown in Figure 6.16. As can be seen, the and Diederichs (2006). The number of data points in the
bulk of the data falls within the GSI range of 40 to 80. various ranges is shown.
Hoek and Diederichs (2006) note that linear or
exponential relationships such as Equation 6.1,
Equation 6.2 or Equation 6.3 cannot adequately
predict modulus of high quality rock masses. They
therefore derived the sigmoidal relationship given in
Equation 6.4 as an upper bound to their data.

EQUATION 6.4

They also showed Equation 6.4 fits the reliable data


from Bieniawski (1978); Serafim and Pereira (1983);
Stephens and Banks (1989) quite well (see Figure 6.17).
The complete version of Equation 6.4 as given
by Hoek and Diederichs (2006) and reproduced
as Equation 6.5, includes a subjective term called
the disturbance factor, D, which ranges from 0 FIGURE 6.17 Comparison of Equation 6.4 with the rock
mass modulus data from various sources as shown as
(undisturbed) to 1 (fully disturbed). The data shown published in Hoek and Diederichs (2006).
in Figure 6.17 corresponds to a D = 0 (undisturbed).

Douglas (2002) collated published data on


measured rock mass modulus, E m, as listed in Table
6.5. Figure 6.18 shows the large scatter in measured
Em and recorded GSI as well as the poor correlation
EQUATION 6.5
between the two.
Superimposed onto Figure 6.18 is the curve suggested
It is important to note that Equation 6.5 was derived
by Hoek and Diederichs (2006), i.e. Equation 6.5,
by curve fitting data. Some of that data that was used
for D = 0 and D = 0.5. It is apparent that the data is
to fit the second form of the equation, the one with
largely bounded by these two curves though neither is
the ratio E m/Ei ,was obtained by Hoek and Diederichs
particularly well-conditioned to the data.
(2006) by assuming the intact modulus Ei is related
to the σc by the typical modulus ratio MR for the
particular rock type. In other words, the relationship
6.4.5 Other
Kayabasi et al. (2003) evaluated 57 datasets comprising
in Equation 6.5 defining the ratio E m/Ei is partly based
in situ modulus derived from plate load tests and GSI
on inferred not measured data.

70
6 | Rock Mass Modulus

FIGURE 6.18 In situ modulus versus GSI from by Douglas (2002). The error bars acknowledge the variation of Em in the test
results. The uncertainty in GSI was due both to the variability of the rock masses and the conversion from published RMR
values to GSI.

values that were reported by others from the Deriner


and Ermenel dam sites in Turkey. The data relates to
very strong quartz-diorite and strong limestone and is
shown in Figure 6.19. The degree to which the rock
mass was disturbed is not known.
Kallu et al. (2015) compared published correlations
with plate load tests from 13 measurements from
mines in Nevada, USA, eight measurements from the
Bakhtiary dam site in Iran and five measurements
from the Portugues dam site in Puerto Rico (Stephens
and Banks 1989) (see Section 6.5.2). Figure 6.20 plots
their data with respect to RMR 76. They curve-fitted the
following awkward exponential and log relationship
between in situ rock mass modulus (E m ) and four FIGURE 6.20 In situ modulus versus RMR76.from Kallu et al.
variables: average block size or joint spacing (BS), field (2015)
estimated rock strength (SF), fracture roughness (RF)
and infill hardness (IF). The reasonable fit achieved is
shown in Figure 6.21.

E m=e-0.731+0.0846SF+0.382lnBS+0.134RF+0.1571F

FIGURE 6.19 Plate load test results of in situ modulus FIGURE 6.21 Comparison of measured and predicted in situ
against GSI from (Kayabasi et al. 2003) modulus from Kallu et al. (2015)

71
ESTIMATING ROCK MASS PROPERTIES

TABLE 6.5 In situ modulus collated by Douglas (2002)


Rock GSI Em [GPa] Testing Source
mean/range mean/range
Claystone 34 30-63 1.6 Pile Loading Santiago (1986)
Mudstone 49 40-49 1.8 Dilatometer Georgiadis (1986)
Claystone 53 44-62 4.4 0.51-8.28 Plate Bearing Gifford (1986)
Siltstone 39 23-60 0.11 0.05-0.17 Pressuremeter Pavlakis (1986)
Sandstone 67 54-79 0.3 0.25-0.36 Plate Loading Poulton (1986)
Basalt 60 51-79 25 10-40 Jointed Block Schultz (1995)
Sandstone 75 55-87 3.9 2.2-5.6 Flat jack Cheng (1993)
Sandstone 75 55-87 3.7 2.3-5.1 Plate Loading Cheng (1993)
Limestone 79 63-91 37.5 32.5-42.5 Plate Loading Giovanni (1993)
Basalt 39 30-74 1.39 0.2-4.6 Pressuremeter Mcdonald (1993)
Granodiorite 52 3.8 Goodman Jack Littlechild (2000)
Granite 55 2.5 Goodman Jack Littlechild (2000)
Siltstone 60 14 Goodman Jack Littlechild (2000)
Tuff 75 11 Goodman Jack Littlechild (2000)
Marble 76 18 Goodman Jack Littlechild (2000)
Sandstone 46 35-57 7.5 5-10 Crosshole Littlechild (2000)
geophysics
Granite 48 41-55 17.5 15-20 Crosshole Littlechild (2000)
geophysics
Siltstone 68 64-71 25 20-30 Crosshole Littlechild (2000)
geophysics
Marble 78 71-84 15 10-20 Crosshole Littlechild (2000)
geophysics
Sandstone 60 55-65 6.6 MPBX Bieniawski (1990)
Shale 63 58-68 13.8 MPBX Bieniawski (1990)
Shale 72 67-77 13.8 MPBX Bieniawski (1990)
Mudstone 75 70-80 5 MPBX Bieniawski (1990)
Shale 75 70-80 12.4 MPBX Bieniawski (1990)
Granite 35 30-40 4 Plate Loading Ribacchi (1984)
Granite 42 37-47 8 Plate Loading Ribacchi (1984)
Granite 43 38-48 10 Plate Loading Ribacchi (1984)
Granite 45 40-50 7 Plate Loading Ribacchi (1984)
Granite 62 58-68 9 Plate Loading Ribacchi (1984)
Granite 64 59-69 8 Plate Loading Ribacchi (1984)
Granite 50 45-55 13 Plate Loading Ribacchi (1984)
Granite 55 50-60 13 Plate Loading Ribacchi (1984)
Granite 57 52-62 18 Plate Loading Ribacchi (1984)
Granite 67 62-72 24 Plate Loading Ribacchi (1984)
Granite 70 65-75 20 Plate Loading Ribacchi (1984)
Granite 75 70-80 13 Plate Loading Ribacchi (1984)
Granite 75 70-80 15 Plate Loading Ribacchi (1984)
Granite 80 75-85 33 Plate Loading Ribacchi (1984)
Granite 80 75-85 35 Plate Loading Ribacchi (1984)

72
6 | Rock Mass Modulus

6.4.6 Collated data Two things are apparent in Figure 6.22.


Figure 6.22 plots the rock mass moduli database • The data is generally in keeping with the range and
collated from the published data presented in distribution given in Hoek and Diederichs (2006)
the preceding sections. The base figure shows the • The curves generated with Equation 6.5 for D = 0,
Chinese and Taiwanese in situ moduli from Hoek 0.5 and 1 do cover most of the data.
and Diederichs (2006). Borehole test data were Figure 6.23 re-plots the data but excludes that
excluded from the Stephens and Banks (1989) and which is likely to represent disturbed rock mass.
Douglas (2002) datasets where known so as to select Much of the undisturbed rock mass data falls within
the most reliable rock mass moduli values (refer to a range about the value predicted by Equation 6.5 for
the Introduction of Section 6.3). The data includes D = 0. The range shown in Figure 6.23 by the dashed
disturbed and undisturbed rock masses. As the lines is generated by replacing the constant 75 in the
published sources did not always include raw data, numerator of the exponential in Equation 6.5 with 70
there may be some duplication in the data presented and 80, respectively, viz
in Figure 6.22. Superimposed are the curves generated
from Equation 6.5 for D = 0, 0.5 and 1 (Hoek and
Diederichs 2006).
EQUATION 6.6

FIGURE 6.22 Collated in situ modulus versus GSI. The base data (grey circles) are from Hoek and Diederichs (2006).

73
ESTIMATING ROCK MASS PROPERTIES

100

Bieniawski (1978)
D = 0

80 Serafim & Pereira (1983)

Stephen & Banks (1989)

Palmstrom & Singh (2001), D=0

60 Kallu (2015)
Em [GPa]

Stripa

40

20

0
0 20 40 60 80 100

GSI

FIGURE 6.23 Range of in situ modulus versus GSI showing the typical bounds for undisturbed rock masses

Also the data from Palmström and Singh (2001) 6.4.7 Conclusion
and Nejati et al. (2014) suggests a value of D = 0.5 The empirical relationships between rock mass quality,
may be appropriate for rock masses disturbed by drill be that RQD, RMR, Q, GSI or other measure, is an
and blast and stress relief. Figure 6.24 presents the exercise in curve fitting data that has been appropriately
data from these two sources together with the curves selected to be representative of that rock mass.
generated from Equation 6.6 with D = 0.5. Data from From the data and discussion presented, it is
other sources is not shown as the extent to which the apparent that a sigmoidal curve, such as that proposed
rock mass from these sources is disturbed is not known. by Hoek and Diederichs (2006) is better suited to the

20
100

18

16
80

14

60 12
Palmstrom & Singh (2001)
Palmstrom & Singh (2001)
Em [GPa]
Em [GPa]

D = 0.5 Nejati et al (2014)


10
Nejati et al (2014)

40 8

20 4

0 0
0 20 40 60 80 100 0 20 40 60 80 100
GSI GSI

FIGURE 6.24 Left - range of in situ modulus versus GSI for drill & blast and stress relief disturbed rock masses. The right
graph is just a close-up of the data.

74
6 | Rock Mass Modulus

data than a straight line or exponential curve.


Further, the data collated suggests that while the
equation proposed by Hoek and Diederichs (2006) is
a good fit to the mean, it is better to acknowledge the
uncertainty in the prediction by quoting the range, i.e.
mean

FIGURE 6.25 Concept of equivalent elastic continuum


(after Kulhawy 1978)
with the range, from

to EQUATION 6.9

Amadei (1996) considers these analytical models


can achieve acceptable predictions of rock behaviour
EQUATION 6.7 from an engineering point of view despite their
limitations. This is particularly so if the selected rock
Furthermore, a value of D = 0.5 appears to be properties are determined in a stress range comparable
appropriate for rock masses disturbed by drill and blast to what is expected in situ. Hence, the analytical
and stress relief. models are appealing though there are several implicit
assumptions in these models.
6.5 EQUIVALENT CONTINUA • The normal and shear stiffness are constant and
independent of stress
There are several analytical models which seek to
relate a jointed mass to an equivalent continuum by • The joints have negligible thickness and hence
following the same basic principle: the deformation of undergo equal strains in their plane
the equivalent continuum is the sum of the deformation • The intact rock and the joints undergo equal strains.
of the intact rock and the deformation of the joints. These assumptions may over-simplify the case.
The model shown in Figure 6.25 is an example of a Consider the rock mass in Figure 6.25 subject to a
relatively simple, theoretical model for a sedimentary vertical load. The horizontal joints, which are spaced at
rock mass. The rock mass can be thought of as blocks of sz, will compress before the intact blocks of rock. The
intact rock separated by evenly spaced, parallel bedding implication is that the modulus of the rock mass changes
planes and orthogonal joints. For the particular model
shown in Figure 6.25, Kulhawy (1978) quotes earlier
work by Goodman et al. (1968) and Duncan and
Goodman (1968) to calculate the rock mass modulus
(Em) as a function of defect spacing for loading
perpendicular to the defects as in Equation 6.8.

EQUATION 6.8

Complexity can be built into this type of model


to include orthorhombic layers (Gerrard 1982); joint FIGURE 6.26 Theoretical behaviour of a defect under
slip (Adhikary and Dyskin 1997; 1998); and/or defect compression, tension and shear loading (Hudson and
shear stiffness (Zhang 2010): Harrison 1997)

75
ESTIMATING ROCK MASS PROPERTIES

under load: it is governed by the stiffness of the joints at Hudson (1980) extended the work of Bristow to
low stress levels and that of the intact rock modulus at account for crack to crack interactions and gave an
high stress. This was observed by Bandis et al. (1981) expression for the shear modulus of a fluid-filled,
and Hudson and Harrison (1997), see Figure 6.26. cracked solid, G:

6.6 COMPLIANT INCLUSIONS


Another way that researchers have looked at analytical
models to estimate mass modulus has been to describe
EQUATION 6.12
fractures, cracks and pores as compliant inclusions in
a stiff, un-fractured solid. This research can be loosely
divided as coming from two streams; geophysics and where K i is the bulk modulus of the uncracked solid =
‘traditional’ rock mechanics. Ei/[3(1−2vi)].

6.6.1 Geophysics As Walsh (1965) mathematically showed that


Appreciation that low-frequency, long-wavelength, though Poisson’s Ratio (v) varied with the crack
seismic waves cannot “see” individual cracks in a rock density and with stress, its variation is within relatively
mass but are governed by overall characteristics of the tight bounds, e.g. 0.15 to 0.30, it is possible to write the
fractured rock mass, suggested that a seismic response approximation v ≈ v I and hence E/Ei ≈ G/Gi.
of a fractured rock could be simulated in a properly Though Hudson’s equation (Equation 6.12) is
selected homogeneous solid. Two important papers actually for the motion of waves through solids and
whose aim was just that – to replace a cracked medium hence dynamic modulus is implied, the values can be
with an homogeneous one that has the same overall taken as static modulus when the wavelength (λ=2π/k)
elastic properties – are Bristow (1960) and Hudson is long compared with the scale-length of the crack (ka
(1980). <<1) (Hudson 1980). This is the case for seismic waves
Bristow (1960) derived equations for the effective through rock masses, as the typical wavelengths for
elasticity of solids that contain a few dry, randomly P- and S-waves are 50 to 500 km and those of surface
oriented, circular, thin cracks. He assumed interactions waves 30 to 1000 km (Mooney 2017), compared to
between cracks could be ignored and that random crack cracks of tens of metres and hence, the constraint for ka
orientations result in overall isotropy. His equation for <<1 as stated by Hudson (1980) is met, i.e. ka ≈10-4 /m.
fractional change in elastic and shear modulus is: Grechka and Kachanov (2006) critique various
analytical expressions for deriving the elastic
properties of cracked solids. They conclude that
expressions such as Equation 6.10 of Bristow (1960)
and Equation 6.12 of Hudson (1980) were “sufficiently
accurate” to estimate elastic properties for solids with
“ irregular and intersecting, approximately flat cracks.”
Grechka and Kachanov (2006) further conclude that
these expressions give acceptable elastic properties
EQUATION 6.10 of fractured rocks for seismic needs, even when the
fractures intersect and form interconnected networks.
where
6.6.2 Rock mechanics
Gi is the isotropic shear modulus of the uncracked Measurement of the deformation of rocks under
solid = Ei/[2(1+vi)] hydrostatic pressure shows that compressibility often
vi is Poisson’s ratio of the uncracked solid is not a constant, as would be expected for an elastic
material, but varies with pressure and with the type of
test (Walsh 1965). The graphs of stress versus strain
and
produced by Hudson and Harrison (1997) show that
theoretical behaviour (Figure 3.4).
EQUATION 6.11
The applicability of theories that describe that
observation by relating the elastic modulus of micro-
cracked materials to the number and size of cracks was
a is the radius of the average crack
explored by Zimmerman (1985). This also included the
n is the number-density of cracks (= number of theory from Walsh (1965) as given in Equation 6.13,
cracks / volume) which is the same equation as given by Bristow (1960)
(Equation 6.10).

76
6 | Rock Mass Modulus

expressions for the effective modulus of isotropic


solids containing randomly oriented spheroids. They
also present equations for Poisson’s ratio (v) and
EQUATION 6.13 bulk modulus (K) for infinitely thin cracks based
on crack density, Γ (Equation 6.14). See Figure 6.27
The starting point is the compressibility of a pore, for example.
which is found by calculating the fractional volume David and Zimmerman (2011) define infinitely thin
change of a single cavity subjected to a hydrostatic cracks as having aspect ratios, α ≤ 0.01 though note
pressure over its surface. Zimmerman (1985) concluded that the error is less than 2% compared with the result
the theories adequately predict the change in elastic of more rigorous theoretical calculations when using
modulus with increased micro-cracking. Equation 6.14 with α up to 0.1.
Zimmerman (1991) further developed equations
for the compressibility of three-dimensional spheroids
from the compressibility of two-dimensional, circular
and elliptical pores (Walsh et al. 1965) and of thin
cracks (Berry 1960). With each of these approaches
the total pore volume change is implicitly assumed to
be the sum of the volume changes calculated for the
and noting that E = 3K(1-2v), then
individual isolated pores or spheroids. However, in an
actual porous rock, the stress field to which a given
pore is subjected will be influenced by the presence
of nearby pores. Hence, its volumetric strain will be
different to that calculated by ignoring the presence of EQUATION 6.14
other pores [Chapter 11 in Zimmerman (1991)].
Hu and McMechan (2009) showed that these Vernik and Kachanov (2012) also confirm the
theories can predict elastic properties of laboratory test limit of α ≤ 0.01 to be appropriate to assess the elastic
data quite well at small crack densities, i.e. small values properties of dry or fully-drained saturated rocks.
of Γ. Saenger et al. (2004) show that the crack density
threshold is approximately Γ ≤ 0.2, although Grechka 6.6.3 Application
and Kachanov (2006) use a limit of Γ ≤ 0.1. The equations of David and Zimmerman (2011) may
David and Zimmerman (2011) developed the be useful to estimate rock mass modulus if bedding
approach further and derived approximate analytical planes, shears, joints, etc. could be considered as
infinitely thin cracks.
Adopting the limiting aspect ratio which defines
infinitely thin cracks as α ≤ 0.01, in accordance with
David and Zimmerman (2011), implies a 1 m long
defect would have a maximum thickness of 10 mm
(10/1000 = 0.01). Similarly, a 50 m long defect would
have a maximum thickness of 500 mm (500/50000
= 0.01). These limits are not unreasonable for most
rock masses. Further, if we accept the work of David
and Zimmerman (2011) and of Vernik and Kachanov
(2012) that the methodology is still valid and with
small error for the higher limit of α ≤ 0.10, the
maximum thickness increases by a factor of 10.
However, a link between the degree to which a
rock mass is jointed and crack density (Γ) is needed
before the compliant inclusion models can be used to
estimate rock mass modulus. At the scale of interest
to tunnelling, the volumetric joint count ( Jv), and by
extension, block volume (Vb) and RQD may be suitable.
Volumetric joint count ( Jv) is defined as the number
of joints intersecting a cubic metre (Palmström 2005).
This is the same definition as that used for the term n
FIGURE 6.27 Normalised effective bulk modulus of a solid (number of cracks / volume) in Equation 6.11 which is
containing spheroids after David and Zimmerman (2011) used to estimate Γ = na3.

77
ESTIMATING ROCK MASS PROPERTIES

If the jointing occurs mainly in sets, then:

where s1, s2, s3 … sn are the average spacings of the


joint sets.
Palmström (2005) discusses the poor ability of RQD
to characterise the degree of jointing but does give the
following correlation between RQD and Jv (Figure
6.28) which in terms of the scale for most tunnels in
blocky rock, i.e. joint spacing in the order of 0.1 to 1 m
(Figure 4.13), is probably suitable.

Substituting this equation for Jv into Equation 6.11


gives the following approximation.

FIGURE 6.28 Correlation between RQD and Jv given by


EQUATION 6.15 Palmström (2005)

FIGURE 6.29 Comparison of the range covered by RQD with those of joint spacing, block volume and volumetric joint count;
from Palmström (2005)

78
6 | Rock Mass Modulus

6.7 ILLUSTRATIVE CASES Hawkesbury Sandstone. Two sub-vertical joint sets


occur and are typically planar, rough & tight. One set
Distinct rock masses – anisotropic sedimentary
rock masses and an isotropic igneous rock mass – spaced at approximately 3 m, the other at 5 to 10 m.
which have substantial testing databases are used in This spacing of defects implies typical block volumes
this section as case studies to illustrate the various of between 15 to 30 m3, ranging up to 150 m3.
approaches discussed in this chapter to estimate rock A five-group classification system for foundation
mass modulus. The Ashfield Shale and Hawkesbury design on the sandstones and shales in Sydney was
Sandstone are found in Sydney, Australia. The Lac du developed by Pells et al. (1978) and updated 20 years
Bonnet Granite is found in the AECL’s Underground later in Pells et al. (1998). The classification system
Research Laboratory in Manitoba, Canada. Typical is based on intact rock strength, defect spacing and
parameters for the rock masses are given in Table 6.6. allowable seams (clay, fragmented, highly weathered
or similar zones). For example, Class II Hawkesbury
6.7.1 Ashfield Shale & Hawkesbury Sandstone has an unconfined compressive strength,
Sandstone σc > 12 MPa; defect spacing > 600mm and <3%
The Ashfield Shale has a maximum thickness of allowable seams.
approximately 60 m and grades from a claystone, There is substantial published data regarding the
siltstone to a siltstone / fine grained sandstone laminate properties of Hawkesbury Sandstone and Ashfield
(MacGregor 1985). It has been extensively worked Shale; e.g. Bertuzzi (2014); Bertuzzi (2014); Bertuzzi
throughout Sydney as a source of brick-making clay and Pells (2002); Clarke and Pells (2004); Pells (2004);
(Branagan and Packham 2000). The Ashfield Shale Pells et al. (1998); Won (1985). The typical rock mass’
comprises well developed, thinly laminated rock Young’s modulus for vertical loading is provided in
with wider spaced bedding from about 0.1 - 0.5 m to Pells et al. (1998). For Hawkesbury Sandstone for
2m. The bedding planes are typically sub-horizontal, example, the values are: Class I > 2000 MPa; Class II
undulating and slightly rough with occasional silty 900 to 2000 MPa; Class III 350 to 1200 MPa; Class
clay infill. Two sub-vertical and orthogonal joint sets IV 100 to 700 MPa; and Class V 50 to 100 MPa. The
occur but so too do other joints of variable orientation. Young’s modulus in the horizontal direction for Class
Joints are typically planar, tight and slightly rough II sandstone is given as 1500 MPa in Pells et al. (1994).
(Bertuzzi 2014; Bertuzzi 2014). A small percentage The normal stiffness of the typical bedding planes
of swelling clays is present in Ashfield Shale though can be estimated using the conclusions of Chapter 3.3.
generally does not pose an issue in the rock mass. For example, bedding planes in either Ashfield Shale
Soils derived from Ashfield Shale are slightly to or Hawkesbury Sandstone occasionally have clayey /
moderately expansive with measured swells of 2% to 7% silty / sandy infill. This infill can be assumed to have
(Won 1985). an undrained shear strength su = 50 kPa and modulus
The Hawkesbury Sandstone is a medium to coarse E = 20 MPa, which suggests an equivalent normal
grained, quartzose sandstone which was deposited in 1 stiffness of k n = 1.6 x 20 MPa / [0.005 to 0.010] =
to 5 m thick beds, typically 2 m, that are either massive 3200 to 6400 MPa/m for typical bedding planes.
or cross-bedded (Figure 6.32 and Figure 6.33). Shale This is in the range obtained in the laboratory testing
breccia is common at the contacts between individual (see Chapter 3.3) and measured by Clarke and Pells
beds, with siltstone interbeds forming a minor part (2004). Assuming σn > 1 MPa, the shear stiffness of
of the overall Hawkesbury Sandstone unit. Cross- the bedding plane (k s) can be estimated using the ratio
bedded beds make up approximately 70% of the k s / k n = 0.10 which implies k s ≈ 320 to 640 MPa/m.

79
ESTIMATING ROCK MASS PROPERTIES

FIGURE 6.30 Typical photos of Ashfield Shale Class II borehole core (Bertuzzi 2014)

FIGURE 6.31 Photo of typical Ashfield Shale Class II exposed in the Castle Hill Crossover Cavern.

80
6 | Rock Mass Modulus

FIGURE 6.32 Typical photos of Hawkesbury Sandstone Class II borehole core (Bertuzzi 2014)

FIGURE 6.33 Photo of saw cut Hawkesbury Sandstone Class III becoming Class II beneath an existing house

81
ESTIMATING ROCK MASS PROPERTIES

6.7.2 Lac du Bonnet Granite Limited (AECL) Underground Research Laboratory


The Lac du Bonnet granite is a very high strength in Manitoba, Canada encountered six fractures, each
(σc = 170 to 220 MPa), relatively undifferentiated with a trace length less than 1.5 m. The implied block
and homogeneous, pink and grey, medium to coarse- volume is therefore in the order of thousands of cubic
grained, massive, un-jointed, porphyritic granite- metres, say greater than 1000 m3.
granodiorite (Martin et al. 1997), see Figure 6.34. The Also, it is reasonable to assume that the average
data regarding its properties is provided in Diederichs sheared joint length will be less than a metre to a
(2007); Diederichs et al. (2004); Martin (1990); maximum of a few metres. Using the conclusions
Martin (1993); Martin et al. (1999); Martin et al. of Chapter 3.3 this implies stiffness of k s ≈ 1000 to
(1997). Martin et al. (1999) quote RMR ≈ 100 and 10,000 MPa/m and k n ≈ 10,000 to 100,000 MPa/m.
Diederichs (2007) gives mi= 30, s = 1 and a = 0.5 for In terms of rock mass modulus, Martin (1990)
intact and GSI = 85, mb= 17.6, s = 0.19 and a = 0.5 for concluded that while anisotropy is evident in core
Lac du Bonnet granite rock mass. samples of Lac du Bonnet granite, there is no evidence
The Lac du Bonnet granite is a massive rock mass which suggests the in situ granite is an anisotropic rock
with very few joints. Martin et al. (1997) record that mass. The small scale anisotropy is considered to be
approximately 500 m length of tunnel through Lac a function of stress-relief micro-cracking related to
du Bonnet Granite at the Atomic Energy of Canada drilling and sampling (Martin 1990).

FIGURE 6.34 Photo of the Lac du Bonnet granite in the Mine-By Experiment, Underground Research Laboratory, AECL,
Manitoba, Canada http://www.civil.engineering.utoronto.ca/Page611.aspx accessed 13 September 2014

82
6 | Rock Mass Modulus

TABLE 6.6 Typical Parameters

TYPICAL PROPERTIES HAWKESBURY SST ASHFIELD SHALE LAC du BONNET


CLASS II CLASS II GRANITE

12 – 50 MPa 7 – 40 MPa 155 – 248 MPa


Intact rock UCS
average ≈ 25 MPa average ≈ 15 MPa average ≈ 200 MPa

5 – 20 GPa 5 – 20 GPa 50 – 80 GPa


Intact modulus, Ei
average ≈ 10 GPa average ≈ 10 GPa average ≈ 65 GPa

Poisson ratio, v 0.25 0.25 0.26 – 0.3

1.1 (saturated)
Anisotropy ratio (saturated) ≈1
1.3 (oven dried)

Rock mass modulus, Vertical 0.9 – 2.0 GPa


Em Vertical 0.7 – 2.0 GPa Typically, 60 GPa
Horizontal 1.5 GPa

Em / Ei. ≈ 0.1 – 0.3 ≈ 0.1 – 0.3 ≈ 0.9 – 0.95

RQD 90 – 100% 80 – 100% 100%


Classification

RMR’89 70 – 85 60 – 75 90 – 100

GSI 55 – 75 45 – 55 80 – 95, 85

Q 5 – 75 1.5 – 50 ≥ 200

Spacing: 0.5 to 2 m,
Spacing: 1 to 5 m, 2 m Length: > 10 m
Length: > 10 m Shape: Planar to NA
Bedding Shape: Undulating, rough undulating, smooth Massive
with occasional ≤ 10 mm to slightly rough with
clayey silty sand occasional clay seams up to
10 mm thick

Spacing: 1 to 2 m Spacing: > 10 m


Spacing: 3 m; 5 to 10 m
Length: within beds Length: < 1.5 m
Jointing Length: within beds
Shape: planar, smooth to Shape: undulating, very
Shape: planar, rough, tight
slightly rough, tight rough, tight

Block volume 15 to 30 m3 0.5 to 5 m3 > 1000 m3

Parameter Bedding Bedding Massive

In situ In situ In situ


Scale Lab Lab Lab
(20 m) (20 m) (5 m)
Main defects

JRC 4 to 6 2.5 4 to 6 2.5 8 to 12 3

JCS
25 10 20 7 200 100
MPa

φr 30° 25° 30°

83
ESTIMATING ROCK MASS PROPERTIES

6.7.3 Classification systems


6.7.3.1 RQD
The mean correlation between RQD and Em / Ei as
presented in Section 6.5.1 and shown in Figure 6.35
suggests for our two illustrative cases:
• Em / Ei = 0.38 to 0.89 for Ashfield Shale with RQD
= 80 to 100%
• m / Ei = 0.58 to 0.89 for Hawkesbury Sandstone
E

with RQD = 90 to 100%
• Em / Ei = 0.89 for the Lac du Bonnet Granite with
RQD = 100%.
This seems to be a good predictor for the Lac du
Bonnet Granite (Em / Ei ≈ 0.9 to 0.95) but a poor
predictor for the Ashfield Shale and Hawkesbury
Sandstone (Em / Ei ≈ 0.1 to 0.3).
6.7.3.2 RMR
Using Figure 6.36 and Bieniawski’s Equation 6.1,
which is Em = 2RMR – 100 and Serafim & Pereira’s
Equation 3 , which is Em = 10 ^[(RMR-10)/40]
• Ashfield Shale with an RMR76 ≈ 55 – 70 suggests FIGURE 6.35 Correlation between RQD and Em/Er, after
Zhang (2010); Zhang and Einstein (2004). Lines are drawn
values of Em = 15 – 40 GPa
(dot-dashed, orange) at RQD=90% and 100% which gives
• awkesbury Sandstone with an RMR76 ≈ 65 – 80,
H
 mean Em/Ei = 0.58 and 0.89, respectively.
Em = 25 – 60 GPa
• Lac du Bonnet granite with an RMR76 = 90 – 100,
Em = 80 to > 100 GPa.
None are particularly good matches with the
actual data.

FIGURE 6.36 Correlation between RMR and Em. Lines are drawn for Hawkesbury Sandstone (dot-dashed, orange) and Luc du
Bonnet Granite Sandstone (dot-dashed, blue).

84
6 | Rock Mass Modulus

The Mohammadi and Rahmannejad (2010) 100

equation [E m = 0.0003RMR 3 – 0.0193RMR 2 +


0.3157RMR + 3.4064] suggests: D = 0

• Em = 12 – 34 GPa for Ashfield Shale with an RMR76 ≈ 80

55 – 70 71

• Em = 25 – 59 GPa for Hawkesbury Sandstone with


an RMR76 ≈ 65 – 80 60

• Em = 94 – 142 GPa for Lac du Bonnet granite with

Em [GPa]
50

RMR76 = 90 – 100.
This correlation between RMR and modulus also 40

tends to overestimate the mass modulus.


6.7.3.3 Q
20
The relationship between Em and Q given Equation 14

6.3, which is M = 10Qc1/3, suggests:


• Em=5 to 37 GPa for Ashfield Shale where Q = 1.5 0
– 50 0 20 40 60 80 100

• Em=8 to 33 GPa for Hawkesbury Sandstone, Q = 5 GSI

– 75
• m 68 GPa for Lac du Bonnet Granite, Q 200.
E FIGURE 6.37 Correlation between GSI and Em. Lines are
The correlation overestimates the rock mass modulus drawn for Hawkesbury Sandstone (dot, orange) and Luc du
Bonnet Granite Sandstone (dot, blue).
for Ashfield Shale and Hawkesbury Sandstone but
predicts it well for Lac du Bonnet granite.
6.7.3.4 GSI
Using a disturbance factor, D = 0 for ‘undisturbed rock The GSI correlation overestimates the rock
mass’, the correlation between Em and GSI (Equation mass modulus for Ashfield Shale and Hawkesbury
6.7) becomes: Sandstone II but provides a good estimate for Lac du
Bonnet Granite.
6.7.3.5 Summary
A comparison between measured Em and that predicted
using the various correlations with classification
with the range, from systems discussed in the preceding sections is
summarised in Table 6.7. The rock mass modulus
obtained using these correlations with RQD, RMR, Q
and GSI is:
• Typically, an order of magnitude larger than that
measured in the field for the medium strength,
sedimentary Ashfield Shale and Hawkesbury
which for: Sandstone.
• Ashfield Shale Class II with GSI = 45 to 55, gives a • The correlation between the predicted and measured
mean Em = 6 to 14 GPa with ranges of 4 – 9 GPa Em for the very high strength, massive Lac du Bonnet
and 9 – 20 GPa, respectively Granite is however better.
• Hawkesbury Sandstone Class II with GSI = 55 to
75, gives an Em = 14 to 50 GPa with ranges of 9 – 39
GPa and 20 – 61 GPa, respectively
• ac du Bonnet Granite with GSI = 85, gives Em = 71
L

with a range of 61 – 80 GPa.

85
ESTIMATING ROCK MASS PROPERTIES

TABLE 6.7 Rock mass modulus predicted from the correlations with UCS, RQD, RMR, Q and GSI

CORRELATION HAWKESBURY ASHFIELD SHALE LAC du BONNET


SANDSTONE CLASS II GRANITE
CLASS II

Em [GPa] Em [GPa] Em [GPa]

Measured Vert 0.9 – 2.0 0.7 – 2.0 65


Horz 1.5

UCS 2.5 – 10 1.4 – 8 31 – 50

RQD 3 – 18 2 – 18 45 – 71

2RMR-100 30 – 60 10 – 40 70 – 90

Figure 6.13 24 – 56 13 – 32 75 – 133


RMR

exp(RMR-100)/36

Mohammadi and 25 – 59 12 – 34 94 -142


Rahmannejad (2010)

10Qc1/3 8 – 33 5 – 27 68 – 107
Q

8Q0.4 15 – 45 9 – 38 67 – 96
(max Q) (max Q)

GSI (mean) 14 – 50 6 – 14 61 – 86

6.7.4 Equivalent continua which gives E m ≈ 3.9 GPa with a range between 1.9
The models presented in Section 6.5 are used to and 14.3 GPa. Alternatively, Equation 6.9 gives typical
estimate the modulus of the equivalent continua as E m of 3.1 GPa again with a range of 1.6 and 12.9 GPa:
given in Equation 6.8, Equation 6.9 and Table 6.8.
For Ashfield Shale Class II, Equation 6.8 suggests that
the rock mass modulus in the vertical direction is:

For Lac du Bonnet Granite, Equation 6.8

which gives E m ≈ 2.4 GPa with a range between 1.2


and 10.0 GPa. Alternatively, Equation 6.9 gives typical
E m of 3.1 GPa again with a range of 0.9 and 8.5 GPa:
gives E m ≈ 58 GPa with a range between 33 and 74
GPa. Equation 6.9 gives typical E m ≈ 56 GPa with a
range of 29 and 72 GPa:

For Hawkesbury Sandstone Class II, Equation 6.8


suggest that the rock mass modulus in the vertical
direction is:
These values are summarised in Table 6.8 along
with the rock mass data. The estimates of rock mass
modulus derived using the equations for an equivalent
continuum are slightly greater than the measured

86
6 | Rock Mass Modulus

TABLE 6.8 Prediction of rock mass modulus using Kulhawy and Zhang equations
CORRELATION HAWKESBURY ASHFIELD SHALE LAC du BONNET
SANDSTONE CLASS II GRANITE
CLASS II

Intact modulus, Ei [GPa] 5 to 20 5 to 20 50 to 80


(Table 6.6) typical 10 typical 10 typical 65

Spacing [m] 1 to 5 0.5 to 2 NA


typical 2 typical 1
Bedding

Normal stiffness, kn 3200 3200 NA


[MPa/m]

Shear stiffness, ks 320 320 NA


[MPa/m]

Spacing 5 to 10 1 to 2 >10
Jointing

typical 3

Normal stiffness, kn 10,000 10,000 10,000 to


[MPa/m] 100,000

Shear stiffness, ks 1000 1000 1000 to


[MPa/m] 10,000

Duncan and Goodman 1.9 to 14.3 1.2 to 10.0 33 to 74


Modules Em [GPA]

(1968) typical 3.9 typical 2.4 typical 58


Rocke Mass

Equation 6.8

Zhang (2010) 1.6 to 12.9 0.9 to 8.5 29 to 72


Equation 6.9 typical 3.1 typical 3.1 typical 56

In situ 0.9 – 2.0 0.7 – 2.0 60


(Table 6.6)

values for Ashfield Shale and Hawkesbury Sandstone Hudson (1980):


but compare well with that for Lac du Bonnet Granite.
It is considered this is a direct result of the lack of
defects in the Lac du Bonnet Granite.

6.7.5 Compliant inclusions


The equations to estimate modulus assuming that
fractures, cracks and pores are compliant inclusions in noting that v ≈ v I and hence E/Ei ≈ G/Gi
a stiff, un-fractured solid are presented in Section 6.6
(Equation 6.10 to Equation 6.14). That is, EQUATION 6.12

Bristow (1960); Walsh (1965): David and Zimmerman (2011):

EQUATION 6.10

EQUATION 6.14

The input data and estimates are developed in


EQUATION 6.13 Table 6.9.

87
ESTIMATING ROCK MASS PROPERTIES

As can be seen in Table 6.9, the crack models of The prediction using the pore model (David and
Bristow, Walsh and Hudson slightly overestimate Zimmerman 2011) is a better match for the Ashfield
the rock mass modulus of the Ashfield Shale and Shale and Hawkesbury Sandstone and slightly under-
Hawkesbury Sandstone but correlate very well predicts that for the Lac du Bonnet Granite.
with the in situ modulus of Lac du Bonnet Granite.

TABLE 6.9 Prediction of rock mass modulus using Hudson and David & Zimmerman equations

PROPERTIES HAWKESBURY ASHFIELD SHALE LAC du BONNET


SANDSTONE CLASS II GRANITE
CLASS II

Ei 5 to 20 5 to 20 50 to 80
(Table 6.6) typical 10 typical 10 typical 65
Intact modulus, [GPa]

Poisson’s ratio, v 0.25 0.25 0.25


(Table 6.6)

Ki = Ei /[3(1-2vi)] 3.3 to 13.3 3.3 to 13.3 3.3 to 53.3


typical 6.7 typical 6.7 typical 43.3

Gi = Ei /[2(1+vi)] 2 to 8 2 to 8 20 to 32
typical 4 typical 4 typical 26

RQD 90 – 100% 80 – 100% 100%


(Table 6.6)
Block size

≤ 4 – 8 joints/m3 ≤ 4 – 12 joints/m3 0.1 – 0.3


Jv
joints/m3

Vb 15 – 30 m3 0.5 – 5 m3 > 1000 m3


(Table 6.6)

Defect length > 10 m Bedding > 10 m Bedding > 1.5 m Jointing


(Table 6.6) (∴ 1 m in 1 m3 (∴ 1 m in 1 m3 (0.3 - 1 m in 1 m3
block) block) block)

Defect thickness ≈ 10 mm ≈ 10 mm Tight (joint)


Crack

(Table 6.6) Bedding Bedding Assume ≤ 1 mm

〈a〉 [m] within 1 m3 0.5 0.5 0.15 – 0.5

Γ =Jv 〈a〉3 0.50 – 1.00 0.50 – 1.50 0.0003 – 0.038

Bristow and Walsh 1.8 - 11 1.4 – 11 50 - 78


Predicted rock mass
modulus, Em [GPa]

(Equation 6.13)

Hudson 1.1 – 11 0.4 – 11 50 – 79


(Equation 6.2)

David & Zimmerman 0.5 – 4.8 0.2 – 4.8 30 – 52


(Equation 6.14)

Rock mass modulus, Em [GPa] 0.9 – 2.0 0.7 – 2.0 60


(Table 6.6)

88
6 | Rock Mass Modulus

6.8 CONCLUSIONS Well-known rock masses – the Ashfield Shale,


Several approaches to estimate the in situ rock mass the Hawkesbury Sandstone and the Lac du Bonnet
modulus, Em, were assessed. In so doing, the inherent Granite – were used as illustrative cases. This chapter
problems with curve-fitting correlations to published showed that the classification systems commonly
data are highlighted. used for tunnel design – Q, GSI (and RMR)– could
provide reasonable estimates of modulus for very high
• The in situ test methodology influences test results, strength, massive rock masses such as the Lac du
e.g. the unreliability of borehole pressuremeter tests
Bonnet Granite.
obtained in very poor quality but high strength rock
However, the classification systems typically over-
• Different in situ test methodologies measure the estimate by an order of magnitude the modulus of the
modulus of different volumes of rock
medium to high strength, sedimentary Ashfield Shale
• The inherent scatter in measured Em means that and Hawkesbury Sandstone, refer to Table 6.7 and
any curve fitting approach will be expected to be a
Figure 6.38.
poor predictor
Analytical models, such as those of Kulhawy (1978)
• The empirical relationship between GSI and Em and Zhang (2010), did provide acceptable predictions
proposed by Hoek and Diederichs (2006) provides
a good fit to the mean. A better approach is to of rock mass modulus for the rock masses: Table 6.8
recognise the uncertainity in the prediction by and Figure 6.38. So too did the compliance models,
quoting a range as given in Equation 6.7. of Hudson (1980) and David and Zimmerman (2011):
Table 6.9 and Figure 6.38. The tunnel designer should
• A value of D = 0.5 for use in Equation 6.7 appears make use of the analytical and compliance models to
to be appropriate for rock masses disturbed by drill
estimate rock mass modulus.
and blast and stress relief.

FIGURE 6.38 Comparison of the various correlations (patterned columns) and of the analytical and compliance models (solid
columns) with measured rock mass modulus for Ashfield Shale (top) and Hawkesbury Sandstone (bottom).

89
ESTIMATING ROCK MASS PROPERTIES

FIGURE 6.39 Comparison of the various correlations (patterned columns) and of the analytical and compliance models (solid
columns) with measured rock mass modulus for Luc du Bonnet Granite.

90
7 | Back-Analysis of Rock Mass Modulus

7 BACK-ANALYSIS OF
ROCK MASS MODULUS

7.1 OUTLINE
Chapter 6 presented and discussed published data
pertaining to rock mass moduli. Most that data stems
from relatively small scale testing – laboratory, borehole
and plate load testing – with some measurements of
convergence of underground excavations. This chapter
presents case studies where the rock mass modulus is
estimated from routine monitoring data obtained
during the construction of underground excavations.
The case studies comprise:
• The widening of the existing twin tunnels of the
M2 motorway in Sydney. These tunnels are within
Hawkesbury Sandstone.
• The excavation of the Northwest Rail Link’s
crossover cavern at Castle Hill, Sydney. The cavern FIGURE 7.1 Circular hole in an infinite medium
is within Ashfield Shale.
The aim of the work presented in this chapter is
twofold: (2005) for the problem shown in Figure 7.1. The
solutions for displacements are given in Equation 7.1
• To develop a relatively simple procedure following in the form quoted in Brady and Brown (2005).
the methodology of Amadei and Savage (1991)
to derive rock mass modulus from the type of Radial
displacement data routinely recorded during typical
tunnelling excavations. While in situ stresses
can and have been back-analysed from detailed
monitoring, such as that carried out at the Atomic
Energy of Canada Limited’s Underground Research
Laboratory (Read 1994), most tunnel projects Tangential
typically do not have such extensive and intensive
monitoring.
• o compare these back-analysed values of rock
T

mass modulus with those values obtained from
the various empirical and theoretical techniques
discussed in Chapter 6.
EQUATION 7.1

7.2 THEORY The displacements of the wall of a circular tunnel


of radius a can be found in terms of Young’s modulus
7.2.1 Isotropy (E) rather than shear modulus (G) by noting that G
Solutions for the stresses and displacements around = E/ 2(1+v) in isotropic, linearly elastic, homogeneous
a circular opening in an isotropic, linearly elastic, media, and that at the tunnel wall r = a. Further, if p
homogeneous medium under plane strain were first represents the in situ vertical stress and if v = 0.25, as
published by Kirsch in 1898. They appear in numerous it is reasonably assumed for most rock masses and as
soil and rock mechanics texts such as Poulos and Davis suggested by the data presented in Gercek (2007), then
(1974), Jaeger and Cook (1979) and Brady and Brown Equation 7.1 becomes:

91
ESTIMATING ROCK MASS PROPERTIES

And hence we can write

EQUATION 7.2
EQUATION 7.3
The choice here to use the apparent modulus (Ea ) in
Equation 7.2 will become evident in the next section. In the general case, numerical analyses can be used
Equation 7.2 is specific to the displacement of the to determine the values of the influence functions,
wall of a circular tunnel. However, it can be generalised I1, I2 , I3 , and I4 as functions of locations around the
to cater for the displacement around a tunnel of any perimeter.
shape by introducing functions to represent the shape
of the excavation (Pells et al. 1980). Re-arranging 7.2.2 Anisotropy
Equation 7.2 and substituting for influence functions, Equation 7.3 assumes isotropic, linearly elastic
I1, I2 , I3 , and I4 the following can be derived. and homogeneous conditions. If the rock mass is
not isotropic and homogeneous, then the modulus
from Equation 7.3 is an apparent modulus, Ea,
in the direction of the measured displacement.
The magnitude of Ea potentially varies along the
tunnel’s circumference.
Amadei and Savage (1991) provide a methodology
using closed-form solutions to determine the elastic
Let properties of anisotropic rock masses from expansion
tests in circular boreholes and tunnels. For transversely
isotropic conditions, they show that the ratio of
apparent modulus (Ea) to Young’s modulus in the
plane of isotropy (E) (which is perpendicular to the
direction of the tunnel) is:

EQUATION 7.4

92
7 | Back-Analysis of Rock Mass Modulus

Where E and E’ are the Young’s moduli (G and G’ 5  e region on the graph that encloses the intersection
Th
are the shear moduli) in the plane of transverse isotropy points of the various K versus Ea lines, gives the rock
and in the direction normal to it, respectively. mass modulus and in situ stress ratio (see example
The typical range for the ratio E/E’ is between 1 and in Figure 7.3.
4 and that for G/G’ is between 1 and 3 (Amadei 1996).
In other words, the stiffness of most rocks is lower in
the direction normal to the major rock planar defects
(Amadei et al. 1987). This is an expected result.
The relationship between Ea /E and E/E’ for the case
when G/G’ = 1 as an example, is shown in Figure 7.2.

FIGURE 7.2 Polar variation of Ea /E with the angle θ for FIGURE 7.3 Example of Ea v K plot
various E/E’ ratios in accordance with Equation 7.4 for the
case G/G’ = 1 and assuming that the Poisson’s ratios, v
and v‘, for rock can both be assumed to be equal to 0.25.
Rounding error is causing the lines to not plot exactly at Ea /E
= 1 at θ = 0°.

7.2.3 Methodology
The methodology used to back-analyse the rock mass
modulus from routine displacement data recorded
during underground excavation is detailed in the
following steps.
7.2.3.1 Isotropic rock mass
The first steps consider the excavation is carried out
within an isotropic rock mass.
1 The influence functions, I1, I2 , I3 , and I4 are defined by
carrying out numerical analyses of the underground
excavation within an isotropic, linearly elastic and
homogeneous medium
2 The linear relationship between Ea, K and u given
in Equation 7.3 is derived for each measured
displacement recorded during underground
excavation
3 The linear relationships are plotted on a graph as K
versus Ea
4 Steps 2 and 3 are repeated for the measured
displacement data from all the monitored points FIGURE 7.4 Example of an arbitrary excavation

93
ESTIMATING ROCK MASS PROPERTIES

7.2.3.2 Anisotropic rock mass The methodology described in the preceding


The next steps consider the excavation is carried out paragraphs requires knowledge of the total displacement
within an anisotropic rock mass. induced by the excavation. However, often the routine
6 The value of θ appropriate for the position of the displacement data measured during tunnelling start
monitoring point as shown in Figure 7.4 is selected. from within the tunnel; that is after the tunnel face
For example, if the monitoring point is the wall at has been excavated.
mid-height, θ =0°; if the monitoring point is the These measurements miss the elastic response of the
centre of the crown then θ =90° (Figure 7.4) rock mass that occurs before the monitoring points are
7 The Ea/E ratios at the relevant θ from Step 6 are installed. Even if the monitoring points were installed
calculated using Equation 7.4 for the typical range at the face, which is usually practically impossible
of anisotropy (Amadei 1996), i.e. E/E’ = 0.7, 1.0, from a physical construction and a personal safety
1.5, 2.0, 3.0 and 4.0 and G/G’ = 1.0, 2.0 and 3.0. perspective, up to 30% of the elastic response of the
For convenience, the ratios are summarised in Table rock mass has occurred ahead of the tunnel face and is
7.1 for θ = 0 and 90° not recorded, e.g. Read (1994).
8 The relationships between Ea, K and u derived in
Step 2 are used for the appropriate monitoring
point position and Ea/E ratio 7.3 CASE STUDIES
The field data collected and collated during
9 The results from Step 8 are plotted on a graph as K
underground excavation works from two recent
versus Ea
projects in Sydney is used to assess the applicability
10  Steps 8 and 9 are repeated for all the measured of the technique developed in the preceding section to
displacement data back-analysing the in situ stress ratio and the moduli of
11 The combination of E/E’ and G/G’ that yields the anisotropic rock masses. The case studies, the locations
tightest concentration of points on the K versus Ea of which are shown in Figure 7.5, comprise:
graph is selected
• The widening of the existing twin tunnels of the
12 The region on the graph that encloses the intersection M2 motorway in Sydney. These tunnels are within
points gives the rock mass modulus and in situ stress Hawkesbury Sandstone.
ratio (Figure 7.3).
• The excavation of the Northwest Rail Link’s
crossover cavern at Castle Hill, Sydney. The cavern
TABLE 7.1 Ea/E values for various E/E’ and G/G’ is within Ashfield Shale.
ratios at θ = 0 and 90°.

Ea/E values

θ=0° θ =90°

G/G’= 1 2 3 1 2 3

E/E’

0.7 0.98 0.68 0.55 1.26 0.86 0.68

1 1.00 0.70 0.56 1.00 0.70 0.56

1.5 1.01 0.70 0.56 0.77 0.55 0.45

2 1.01 0.70 0.56 0.65 0.47 0.38

3 0.98 0.68 0.55 0.52 0.38 0.31

4 0.93 0.65 0.53 0.46 0.33 0.27 FIGURE 7.5 Google map of Sydney showing the location
of the M2 Tunnel and Castle Hill Crossover Cavern

94
7 | Back-Analysis of Rock Mass Modulus

7.3.1 M2 Tunnel Each pair of floor-crown and sidewall convergence


An upgrade of the M2 Motorway in Sydney was data from the 16 locations in the eastbound tunnel
carried out between 2010 and 2013 which included and 14 locations in the westbound tunnel were used
widening the 460 m long twin Norfolk Tunnels in to solve the simultaneous equations derived from the
North Epping from 11.7 m to 15.4 m. The original influence functions in Equation 7.3, i.e.:
tunnels had been excavated wholly within the
Hawkesbury Sandstone between August 1995 and Eastbound
July 1996.
A comprehensive suite of instrumentation was
maintained to provide reliable and robust monitoring
data. This included survey prisms, simple wire
extensometers and tape convergence sections, and
multipoint rod extensometers within each tunnel
(Clarke et al. 2014). The tape convergence sections
included both diagonal and vertical components Westbound
(Figure 7.6). The monitoring programme was more
comprehensive than typical in Sydney because the
tunnels were used by the public motorists each day.
The monitoring data provides 16 locations in the
eastbound tunnel and 14 locations in the westbound
tunnel where both floor-crown and sidewall convergence
was measured. The recorded displacements range:
• Crown sag up to 7 mm
The methodology described in Section 7.3.3 was
• Floor-crown convergence up to 8 mm followed. Interestingly, the introduction of anisotropy
• Wall deflection up to 6 mm did not improve the back-analysed solutions for the rock
Measured crown sag was typically less than mass modulus and in situ stress ratio. The introduction
2 mm. Around Ch 12700 m (the western end of the of anisotropy did not improve the solutions probably
tunnels) the measured crown sag was around 5 mm. due to the variability in actual rock mass modulus and
A comparison of measured crown sag and design the Hawkesbury Sandstone being a low anisotropic
predictions indicate reasonable agreement, with the intact rock following the classification of Singh et al.
design values generally over-predicting displacements. (1989), refer to Table 6.6.

FIGURE 7.6 Monitoring locations shown in the tunnel cross section

95
ESTIMATING ROCK MASS PROPERTIES

Hence, the results obtained assuming an isotropic The values of E and K suggested by the back-analysis
rock mass are summarised in Figure 7.7. The size of are those of the most concentrated grouping of reliable
the points used in the figure represents the subjective points, which are:
classification of data reliability based on magnitude • Eh = Ev = 4500 to 6750 MPa, with an average of say
and trend with respect to time 5500 MPa
• Stable readings with respect to time and • K = 2.7 to 3.1, say 3.0
displacements greater than 1 mm were classified as This compares to the design values of E = 2500 to
‘most reliable’ 10,000 MPa (mean of 6000 MPa) and K = 7.5. While
• Stable readings with displacements less than 1 mm the back-analysed modulus is within the range adopted
were classified as ‘reliable’ in design the back-analysed K is much less. However,
• Unstable trends were classified as ‘unreliable’ this is not surprising as it was found that the measured
• ‘Incomplete’ data was also noted. displacements were less than the design predicted. The
Figure 7.7 shows the modulus and in situ stress ratios presence of the existing twin tunnels, and hence any
inferred from all the data varies between 1100 and stress relief or concentration induced by those tunnels,
25,000 MPa and between 0.1 and 9.8, respectively. was included in the design analysis.
Ignoring the ‘unreliable’ data, these ranges reduce to Highly redundant data is needed given the spread
1100 and 16,000 MPa and between 0.1 and 6.7. shown in this example.

FIGURE 7.7 Results for Eh/Ev=1.0, i.e. assuming isotropic, linear elastic, homogeneous rock mass. Large symbols represent
the ‘most reliable’ data, medium symbols ‘reliable’ data; and the dots the ‘unreliable’ data. Open circles are points from the
Eastbound tunnel; filled circles from the Westbound tunnel. The most concentrated collection of data points is centred about
E = 4500 - 6750 MPa and K = 2.7 - 3.1. The plot on the right is a close-up of the concentrated data.

96
7 | Back-Analysis of Rock Mass Modulus

7.3.2 Castle Hill Crossover Cavern The large span heading was excavated as a split
The Castle Hill crossover cavern, which is excavated in heading. The most reliable data came from prisms
Ashfield Shale, is 160 m long, 23 m wide and 14 to 18 installed within 3.5 to 5 m from the advancing face
m high. Ground cover over the crown ranges from 14 of the second heading. This meant that some of the
to 17 m. The cover / span ratio is approximately 0.65. elastic displacement of the rock mass induced by the
The Castle Hill crossover cavern is one of the largest excavation would not be seen by the prisms. Correction
underground excavations in Sydney, and the largest factors to account for this missed elastic displacement
underground excavation in Ashfield Shale. were assessed and applied to the measured data. A
suite of three-dimensional numerical analyses had
The cavern is part of Sydney’s North West Metro been undertaken as part of the cavern’s design to assess
and allows trains to change tracks. the proportion of elastic displacement that occurred
The monitoring network included surface ahead of the second heading. The design indicated
settlement, inclinometers, extensometers and 16 in- that approximately 60% of the elastic displacement
tunnel convergence arrays which were established as would occur before the prisms are installed in the
the cavern excavation progressed (Figure 7.8). second heading.

FIGURE 7.8 Section overview of cavern monitoring (Shen et al. 2015)

97
ESTIMATING ROCK MASS PROPERTIES

Again, the methodology described in Section 7.3.3 tunnelling excavations. The procedure is simple and
was followed. In this case, introducing anisotropy fast enough to be done during tunnel advance as an
did improve the back-analysed solutions for the rock onsite check of design parameters though highly
mass modulus and in situ stress ratio. The tightest redundant data is needed.
concentration of reliable data points was obtained Two case studies were used:
when Gh /Gv = 1.0 and Eh /Ev = 0.7 (Figure 7.9). The
values of E and K suggested by the back-analysis of the
• The widening of existing twin tunnels of the M2
motorway in Sydney within Hawkesbury Sandstone
crossover cavern excavated within Ashfield Shale are:
• The excavation of the Northwest Rail Link’s
• Eh = 720 MPa, Ev = 1000 MPa crossover cavern at Castle Hill, Sydney within
• K = 1.7 Ashfield Shale.
This compares well with the design values of E = The results of the back-analyses, reproduced below,
300 to 2000 MPa (typically 1000 MPa) and K = 1 to are in keeping with the ranges typically adopted
3.5. The back-analysed anisotropy also compares well in design in Sydney. The results suggest isotropic
with that from laboratory testing, i.e. 1.4 in Table 6.6 conditions for Hawkesbury Sandstone and slightly
from Won (1985). anisotropic conditions for Ashfield Shale.
• Hawkesbury Sandstone
7.4 CONCLUSIONS „„ Eh = E v = 4500 to 6750 MPa, say 5500 MPa
While in situ stresses can and have been back-analysed „„ K = 2.7 to 3.1, say 3.0
from detailed monitoring, such as that carried out at
the Atomic Energy of Canada Limited’s Underground
• Ashfield Shale
„„ Eh = 720 MPa, Ev = 1000 MPa
Research Laboratory (Read 1994), most tunnel
projects typically do not have extensive and intensive „„ K = 1.7.
monitoring. This chapter developed a relatively simple The case studies show that it is possible to back-
procedure that can be followed to derive the in situ analyse the rock mass modulus and in situ stress but
stress ratio and rock mass modulus from the type of highly redundant data is needed given the inherent
displacement data routinely recorded during typical highly variable spread of the monitoring data.

FIGURE 7.9 Results for G/G’=1.0 with E/E’=0.7, 1.0, 1.5 and 2.0

98
8 | Intact Rock Test Data

8 INTACT ROCK
TEST DATA

8.1 OUTLINE „„ Basalt (Brisbane) 23 test results - 23 UCS


In this chapter the data used to compare four intact „„ 
Greywacke (Brisbane) 12 test results - 1 BTS,
strength criteria is presented. The four criteria, which 11 UCS
were identified in Chapter 2 as offering good means of These datasets are used to compare the following
representing intact rock strength, are: four intact strength criteria against that which is most
• The Lade criterion commonly used by tunnel designers – the Hoek-Brown
criterion (H-B).
• The Christensen
• The Hoek-Brown criterion as modified by Mostyn • The Lade criterion (L) (Lade 1977). This criterion
& Douglas was developed from theoretical soil mechanics and
laboratory tests. It is a three-parameter criterion
• The Hoek-Brown criterion as extended to 3D by which has been extended into materials that can
Zhang & Zhu.
sustain small tensile stresses such as rock.
The data comprises:
• The Christensen criterion (C) (Christensen 1997).
• Published test data collated by Mostyn and Douglas This two-parameter theoretical criterion was
(2000). This database is perhaps the most extensive developed for true 3D stresses to cater for a variety
collated to date comprising over 500 data sets of materials.
covering 45 rock types.
• The Hoek-Brown criterion as modified by Mostyn
• Published true 3D testing data collated by and Douglas (2000) (H-B MD).
Colmenares and Zoback (2002) and Kwasniewski
(2013). These two databases are referenced as they • The Hoek-Brown criterion as extended to 3D by
Zhang and Zhu (2007) (H-B 3D).
report results with the influence of the intermediate
principal stress, σ2 . They are also repeatedly These criteria were selected as they have the following
referenced by other researchers. characteristics.
• e artificial laboratory ‘data’ set in Mostyn and
Th
 • A theoretical basis. It is noted here that while the
Douglas (2000), which comprises 58 data points of Hoek-Brown criterion and its derivatives are empirical,
UTS, UCS and TXL. This ‘data’ set is extremely the original work did start with the Griffith crack
useful as it was generated from the assumption that theory. Similarly, the Lade criterion was an empirical
the artificial intact rock complies with the Hoek- development from a theoretical basis.
Brown strength criterion. • A maximum of three parameters. While more
parameters may offer better curve-fitting, such as
• Laboratory test data of intact rock taken from several the You criterion (You 2009), they are more difficult
recent tunnelling projects in Sydney and Brisbane,
Australia. The purpose of this database is to look at to define particularly with often limited laboratory
the ability of the criteria to deal with what is available test data.
in the “real world”. The database comprises: • A smooth, curved failure surface in the octahedral
„„ 
643 test results of Sydney’s Hawkesbury plane.
Sandstone – comprising 185 BTS, 404 UCS and • A curved failure surface in planes containing the
54 TXL hydrostatic axis.
„„ 
Brisbane Tuff, 456 test results - 37 BTS, • Capable of having a tensile strength.
258 UCS and 161 TXL The criteria are also compared with the Modified
„„ 
Phyllite (Brisbane) 166 test results - 8 BTS, 1 Wiebols-Cook criterion (W-C), which was
58 UCS recommended in Colmenares and Zoback (2002) as
„„ 
Quartzite (Brisbane) 79 test results - 4 BTS, the criterion that generally best fitted the compressive
75 UCS failure test data.

99
ESTIMATING ROCK MASS PROPERTIES

8.2 DATA there is “no maximum definable σ1 or where there is


significant doubt as to their accuracy”.
8.2.1 Mostyn & Douglas There are 20 data sets which contain 20 or more
Douglas (2002) compiled a database of over 4500 laboratory tests results. These are indicated in Table 8.1
published laboratory test results of intact rock strength in bold type and comprise 12 sedimentary, 7 igneous
as part of his doctoral thesis. The tests include direct and 1 metamorphic rock types encompassing a UCS
tensile, BTS, UCS and TXL and are collated into 510 range of 2 to 290 MPa.
sets of results comprising 283 sets of sedimentary Out of interest, Table 8.2 shows the best estimate
rocks, 101 of igneous rocks and 126 of metamorphic of the Hoek-Brown strength criterion constants
rocks. The individual rock types contained in the (σc and mi) with those calculated following the Mostyn-
database are listed in Table 8.1. The work was more Douglas approach (σc , mi and a), which allows the ‘a’
broadly published in Mostyn and Douglas (2000). term to be variable instead of fixed at 0.5. The largest
Mostyn and Douglas (2000) argued against improvement in the correlation coefficient made by
excluding results thought to exhibit ductile behaviour following the Mostyn-Douglas approach for these 20
citing work by others that showed that the transition datasets is Δr2 = 0.035, with an average of 0.007. This
between ductile and brittle “ failure is not well defined relatively small improvement in r2 is a function of linear
for all rocks, is often curved and certainly occurs over a regression and masks the improvement in the curve-fit
wide range of stresses.” Hence, the database listed in in the tensile and low confinement zone (Mostyn and
Table 8.1 only excludes published results for which Douglas 2000).

TABLE 8.1 The rock types that are contained within the laboratory test results database compiled by
Douglas (2002). The values in brackets are the numbers of results for each rock type. The 20
datasets which have 20 or more test results are also shown in bold. For example, Sandstone
(1170) – 6 has 1170 test results and 6 datasets of 20 or more tests.
SEDIMENTARY (2387) IGNEOUS (1350) METAMORPHIC (819)
Anhydrite (9) Andesite (29) Amphibolite (22)
Biocalcarenite (8) Aplite (8) Chloritite (8)
Chalk (14) Basalt (16) Dolomite (50)
Claystone (38) Diabase (15) Eclogite (14)
Coal (328) - 1 Diorite (17) Gneiss (92)
Fireclay (4) Dolerite (35) – 1 Greenstone (10)
Limestone (381) – 2 Dunite (47) – 1 Greywacke (12)
Mudstone (145) – 2 Gabbro (33) Marble (190)
Salt (38) Granite (789) – 2 Schist (227)
Sandstone (1170) – 6 Granodiorite (57) – 1 Serpentinite (52) – 1
Shale (157) Lamprophyre (6) Slate (141)
Siltstone (94) – 1 Norite (60)
Peridotite (10)
Pyroclastic (10)
Quartzdiorite (27)
Quartzdolerite (38) – 1
Quartzite (90) – 1
Rhyolite (10)
Syenite (7)
Trachite (5)
Tuff (35)
Whinstone (5)

100
8 | Intact Rock Test Data

TABLE 8.2 The datasets with more than 20 test results, from Douglas (2002).

Hoek-Brown Mostyn & Douglas


(a = 0.5)
UCS
Dataset No. Δr2
[MPa]
σc mi r
2
σc mi α r2

[MPa] [MPa]

Dolerite 26 160.0 194.7 22.6 0.981 230.4 14.3 0.543 0.981 0.001

Dunite 20 70.0 81.3 40.0 0.969 85.4 6.7 0.755 0.997 0.028

Blackingstone 48 179.6 194.5 23.9 0.941 185.2 50.0 0.373 0.951 0.009
Quarry
Granite
Igneous

Westerly 21 225.3 212.7 29.2 0.991 212.5 29.3 0.500 0.991 0.000

Granodiorite Micro 27 193.5 17.6 0.983 239.0 10.1 0.554 0.984 0.001

Quartz- Northumber- 38 286.6 290.5 13.9 0.918 288.4 23.6 0.365 0.919 0.002
dolerite land

Quartzite 48 188.4 204.6 16.4 0.874 188.0 28.8 0.410 0.876 0.002

Serpentinite Oulx 22 193.0 8.1 0.972 151.4 17.6 0.434 0.973 0.001
M

Coal Moura 40 32.7 23.3 20.4 0.817 21.5 41.0 0.396 0.820 0.003

Portland 30 50.6 7.2 0.789 30.5 50.0 0.339 0.794 0.004


medium strong
Limestone
Portland strong 30 59.3 83.1 8.9 0.870 68.2 50.0 0.294 0.886 0.016

48 50.0 59.3 6.3 0.939 57.7 8.9 0.423 0.939 0.000


Mudstone
Sedimentary

Melbourne 20 1.6 1.1 40.0 0.963 1.9 4.2 0.792 0.998 0.035

Keuper 113 7.5 8.2 9.0 0.940 6.8 37.7 0.364 0.950 0.010

South African 55 64.5 60.7 15.7 0.955 71.0 4.0 1.000 0.963 0.009

Derbyshire 33 52.7 54.3 18.7 0.976 52.2 38.0 0.402 0.980 0.004
Sandstone
Pennant 31 196.8 204.0 12.6 0.985 194.4 28.2 0.348 0.990 0.005

Darley Dale 27 80.1 77.5 16.5 0.985 78.7 13.9 0.533 0.985 0.000

Darley Dale 20 79.3 81.0 14.6 0.968 86.9 32.5 0.398 0.979 0.011

Siltstone 64 50.0 53.2 7.8 0.938 54.2 6.2 0.562 0.938 0.000

101
ESTIMATING ROCK MASS PROPERTIES

8.2.2 Artificial to Chatswood Rail Link, Lane Cove Tunnel, CBD


Mostyn and Douglas (2000) simulated a very Metro project and the Northwest Rail Link. The test
comprehensive test program with results generated data comprises 643 results on Hawkesbury Sandstone
for a material fully compliant with the Hoek-Brown - 185 BTS, 404 UCS and 54 TXL. The data is plotted
failure criterion: on a principal stress graph in Figure 8.2.
The laboratory test data from several recent
tunnelling projects in Brisbane, Australia – Inner City
Bypass, Clem7, Airport Link and Legacy Way – have
also been collated. The testing comprises 736 results:
The material’s σc and mi were assumed to be normally
distributed with mean ± standard deviation of σc = 10 • Brisbane Tuff, 456 test results - 37 BTS, 258 UCS
± 2 MPa and mi = 12 ± 2 respectively. Hence, the mean and 161 TXL
strength of the artificial rock is given by: • Phyllite, 166 test results - 8 BTS, 158 UCS
• Quartzite, 79 test results - 4 BTS, 75 UCS
• Basalt, 23 test results - 23 UCS
• Greywacke, 12 test results - 1 BTS, 11 UCS
Fifty-eight results generated were: 10 UTS, 20 The results are plotted on principal stress graphs
UCS and 4 TXL tests at each confining pressure of in the following figures at the same scale for ease of
1, 2, 5, 10, 20, 40 and 80 MPa. The data set, which comparison (Figure 8.3 to Figure 8.7).
has been digitised off the graph presented in Mostyn It is obvious that the laboratory test results for these
and Douglas (2000) is reproduced in Table 8.3 and rocks are limited in number and more importantly in
plotted on a principal stress graph in Figure 8.1. Not type, being mainly restricted to UCS tests. It would be
surprisingly, the correlation is very high, r2 = 0.99. easy to ignore this database as offering ill-conditioned
data to curve-fit the empirical criteria. However, it
8.2.3 Tunnelling Projects is purposefully included in this study to assess the
The laboratory test data from several recent or capability of each of the selected key criteria to provide
proposed tunnelling projects in Sydney, Australia have a failure envelope with the type of limited data often
been collated. The tunnelling projects are: Epping available in tunnelling projects.

TABLE 8.3 Artificial intact rock data set – σ3, σ1 pairs – after Mostyn and Douglas (2000)
-1.9, 0 -1.5, 0 -1, 0 -0.9, 0 -0.85, 0 -0.78, 0
UTS

-0.54, 0 -0.51. 0 -0.47, 0 -0.44, 0

0, 5.7 0, 6.0 0, 6.7 0, 7.8 0, 8.0 0, 8.2

0, 8.8 0, 9.1 0, 9.5 0, 10.1 0, 10.2 0, 10.2


UCS

0, 10.3 0, 10.6 0, 10.8 0, 11.5 0, 11.8 0, 13.5

0, 15.0 0, 15.2

1, 14.2 1, 14.3 1, 16.8 1, 20.5

2, 17.0 2, 17.5 2, 18.0 2, 21.6

5, 27.5 5, 29.4 5, 31.5 5, 33.1


TXL

10, 40 10, 45 10, 47 10, 51

20, 60 20, 68 20, 70 20, 75

40, 92 40, 112 40, 115 40, 118

80, 175 80, 185 80, 187 80, 200

102
8 | Intact Rock Test Data

FIGURE 8.1 Data set from artificial laboratory testing programme, after Mostyn and Douglas (2000)

103
ESTIMATING ROCK MASS PROPERTIES

300

280

260

240

220

200
Major Principal Stress, σ1 [MPa]

180

160

140

120

100

80

60

40

20

0
-20 0 20 40 60 80 100
Minor Principal Stress, σ3 [MPa]

FIGURE 8.2 Data set from the testing of Hawkesbury Sandstone comprises 643 test results programme

104
8 | Intact Rock Test Data

300

280

260

240

220

200

180
Major Principal Stress, σ1 [MPa]

160

140

120

100

80

60

40

20

0
-20 0 20 40 60 80 100

Minor Principal Stress, σ3 [MPa]

FIGURE 8.3 Data set from the testing of Brisbane Tuff comprises 456 test results

105
ESTIMATING ROCK MASS PROPERTIES

300

280

260

240

220

200
σ1 [MPa]

180

160

140

120

100

80

60

40

20

0
-20 0 20 40 60 80 100

σ3 [MPa]
FIGURE 8.4 Data set from the testing of Brisbane phyllite - 166 test results

106
8 | Intact Rock Test Data

300

280

260

240

220

200

180

160
σ1 [MPa]

140

120

100

80

60

40

20

0
-20 0 20 40 60 80 100

σ3 [MPa]

FIGURE 8.5 Data set from the testing of Brisbane quartzite - 79 test results

107
ESTIMATING ROCK MASS PROPERTIES

300

280

260

240

220

200

180
σ1[MPa]

160

140

120

100

80

60

40

20

0
-20 0 20 40 60 80 100

σ3 [MPa]

FIGURE 8.6 Data set from the testing of Brisbane basalt - 23 test results

108
8 | Intact Rock Test Data

300

280

260

240

220

200

180
σ1[MPa]

160

140

120

100

80

60

40

20

0
-20 0 20 40 60 80 100

σ3 [MPa]

FIGURE 8.7 Data set from the testing of Brisbane greywacke - 12 test results

109
ESTIMATING ROCK MASS PROPERTIES

8.2.4 Colmenares & Zoback • the


Modified Wiebols & Cook criterion (Zhou
Colmenares and Zoback (2002) collated and then 1994) or the Modified Lade criterion (Ewy 1999)
investigated the 3D laboratory data of five rock types: gave the best fit to laboratory test results for rocks
with highly σ2 –dependent behaviour. These were
• amphibolite from the KTB site (Figure 8.8); data the three extremely high strength rocks in their
provided to Colmenares and Zoback by Chang
and Haimson database; KTB amphibolite, Dunham dolomite
and Solenhofen limestone which had UCS of 300,
• Dunham dolomite and Solenhofen limestone
375 and 450 MPa, respectively.
(Figure 8.9); digitised by Colmenares and Zoback
from Mogi (1971) • perhaps not surprisingly, the Mohr-Coulomb

and Hoek-Brown criteria fitted the data at least
• Shirahama sandstone and Yuubari shale (Figure equally well for those rocks which behaved
8.10); digitised by Colmenares and Zoback from
largely independent of σ2 . These were the two
Takahashi and Kiode (1989).
high strength rocks in their database; Shirahama
They statistically evaluated the Mohr-Coulomb, sandstone and Yuubari shale, which had a UCS of
Drucker-Prager, Modified Wiebols & Cook, Modified 95 and 120 MPa, respectively.
Lade, Hoek-Brown and Mogi criteria and concluded:

FIGURE 8.8 Data set for KTB amphibolite, from Colmenares and Zoback (2002)

110
8 | Intact Rock Test Data

FIGURE 8.9 Data set for Dunham dolomite (top) and Solenhofen limestone (bottom), from Colmenares and Zoback (2002)

111
ESTIMATING ROCK MASS PROPERTIES

FIGURE 8.10 Data set for Shirahama sandstone (top) and Yuubari shale (bottom), from Colmenares and Zoback (2002)

112
8 | Intact Rock Test Data

8.2.5 Kwasniewski 8.3 RESULTS


The laboratory test results from true triaxial strength of A MS Excel ® spreadsheet with Visual Basic macros
47 data sets of intact rock was collated by Kwasniewski was written to curve-fit the criteria. The spreadsheet
(2013). Twenty nine of these data sets were considered utilised MS Excel® Solver(8) running its Generalised
by Kwasniewski to be reliable which included the five Reduced Gradient algorithm. This algorithm is an
data sets published by Colmenares and Zoback (2002). iterative, non-linear, least squares fitting method that
Hence, there are 24 ‘new’ data sets in Kwasniewski finds variables to minimise the sum of the square of
(2013) (Figure 8.11). These are: errors (Equation 8.1) of smooth nonlinear functions.
• Izumi sandstone (two sets)
• Yamaguchi marble (two sets)
• Mizuho trachyte
• Manazuru andesite EQUATION 8.1
• Inada granite
• Orikabe monzonite
• Chichibu crystalline schist
• Horonai sandstone 8.3.1 Mostyn & Douglas
• Westerly granite (two sets) The data sets which contained at least 20 laboratory
tests results were selected for the comparison. This
• Taga limestone filtered database comprises 12 sedimentary, 7 igneous
• Locharbriggs sandstone and 1 metamorphic rock types. The coefficient of
• SSC muddy siltstone correlation, r2, and the standard error of estimate
• Slask sandstone (SEE) values of the individual curve-fits are listed in
• Rozbark sandstone Table 8.4. The best fitting criteria as assessed by the
• Soignies limestone smallest SEE, and often but not always the highest r2,
• Phra Wihan sandstone (two sets) is shown in bold and the worst fitting is faded.
• Phu Phan sandstone The modified Hoek-Brown criterion following
• Phu Kadung sandstone Mostyn and Douglas (2000) had the smallest SEE
• Maha Sarakham rock salt in all of the 19 data sets. The Christensen criterion
• SAFOD granodiorite. had the poorest fit in 12 of the 19 data sets; the Lade
criterion fitted the data worse in the other seven.
8.2.6 Summary
The database used to compare the key strength criteria
comprises 128 data sets with over 7000 laboratory
test results. It includes sedimentary, igneous and
metamorphic rocks with UCS values ranging from less
than 10 MPa to over 300 MPa.

(8)
Frontline Systems (www.solver.com)

113
ESTIMATING ROCK MASS PROPERTIES

FIGURE 8.11 True 3D triaxial data sets from Kwasniewski (2013)

114
8 | Intact Rock Test Data

FIGURE 8.11 CONTINUED

115
ESTIMATING ROCK MASS PROPERTIES

FIGURE 8.11 CONTINUED

116
8 | Intact Rock Test Data

FIGURE 8.11 CONTINUED

117
ESTIMATING ROCK MASS PROPERTIES

TABLE 8.4 The r2 and SEE [MPa] values obtained from curve-fitting the filtered database from Douglas
(2002)
Rock HB-MD HB HB-3D L W-C C
0.981 0.981 0.981 0.978 0.954 0.594
Dolerite
86.98 88.75 88.75 94.48 136.04 405.43
0.997 0.696 0.696 0.996 0.990 0.321
Dunite
38.21 122.25 123.40 43.60 68.60 575.15
0.193 0.135 0.135 0.151 0.126 0.144
Blackingstone Quarry Granite
116.58 120.66 120.66 194.06 157.42 120.01
0.991 0.991 0.991 0.988 0.952 0.684
Westerly Granite
87.16 87.16 87.16 100.79 199.04 508.94
0.984 0.983 0.983 0.981 0.973 0.882
Granodiorite
74.10 75.73 75.73 78.89 94.88 198.28
0.280 0.231 0.231 0.253 0.220 0.241
Northumberland Quartzdolerite
84.57 87.38 87.38 126.49 110.14 86.79
0.905 0.902 0.902 0.838 0.881 0.707
Quartzite
33.41 34.03 34.03 43.70 37.55 58.85
0.973 0.972 0.972 0.966 0.955 0.954
Serpentinite
64.37 65.61 65.31 73.35 83.51 84.62
0.479 0.456 0.456 0.375 0.361 0.406
Moura Coal
14.85 15.18 15.18 21.04 16.45 15.86

Portland Limestone 1 medium 0.151 0.149 0.149 0.166 0.149 0.051


strong 51.44 51.44 51.44 51.44 51.44 76.38
0.005 0.005 0.005 0.028 0.005 0.006
Portland Limestone 2 strong
33.81 44.00 38.06 64.49 42.10 44.28
0.953 0.952 0.952 0.943 0.947 0.936
Mudstone
5.80 5.82 5.82 6.35 6.10 6.75
0.996 6.75 6.75 0.942 0.994 0.814
Melbourne Mudstone
0.84 2.55 2.56 3.21 0.99 5.75
0.952 0.940 0.940 0.896 0.896 0.908
Kueper Sandstone
3.90 4.38 4.38 5.78 5.77 5.42
0.973 0.963 0.963 0.967 0.973 0.795
South African Sandstone
7.10 8.29 8.29 7.85 7.12 19.54
0.393 0.307 0.307 0.307 0.011 0.321
Derbyshire Sandstone
66.33 70.82 70.82 97.26 84.62 70.09
0.990 0.985 0.985 0.923 0.969 0.833
Pennant Sandstone
9.30 11.22 11.22 25.37 16.09 37.34
0.985 0.985 0.985 0.978 0.968 0.814
1
10.45 10.55 10.55 12.83 15.35 37.22

Darley Dale 0.979 0.968 0.969 0.788 0.920 0.901


2
Sandstone 43.10 53.28 52.40 136.82 84.00 93.47
0.983 0.979 0.979 0.907 0.926 0.909
combined
42.04 47.87 46.80 98.31 87.84 97.31
0.954 0.953 0.953 0.944 0.952 0.909
Siltstone
5.31 5.34 5.34 5.83 5.42 7.44

118
8 | Intact Rock Test Data

TABLE 8.5 The r2 values obtained from curve-fitting the artificial test data from Douglas (2002)
Rock HB-MD HB HB-3D L W-C C
0.991 0.991 0.991 0.989 0.975 0.989
Artificial
4.98 5.06 4.94 9.69 8.37 19.35

8.3.2 Artificial 8.3.4 Colmenares & Zoback


The r2 and SEE from curve-fitting the artificial set of Colmenares and Zoback (2002) assessed the 3D
58 laboratory tests in Mostyn and Douglas (2000) laboratory data of five rock types – KTB amphibolite,
is shown in Table 8.5. The three versions of the Dunham dolomite and Solenhofen limestone,
Hoek-Brown criterion gave the same r2 value but the Shirahama sandstone and Yuubari shale. They
modifications of Mostyn and Douglas (2000) and chose the best-fitting combination of parameters
Zhang and Zhu (2007) produced smaller SEE. The for several strength criteria by minimising the mean
Christensen criterion had the poorest fit to the data. standard deviation misfit to the test data; presumably
minimising the following equation.
8.3.3 Tunnelling Projects
The r2 and SEE from curve-fitting the laboratory test
data from recent tunnelling projects in Sydney and
The results are summarised in Table 8.7 for the
Brisbane is shown in Table 8.6. The modified Hoek-
three criteria that best represented the laboratory data
Brown criterion following Mostyn and Douglas (2000)
- modified Wiebols-Cook, modified Lade and Hoek-
had the smallest SEE, and the Christensen criterion
Brown criteria. The criterion with the lowest misfit
the largest, for the Hawkesbury Sandstone data set. No
(best correlation) is shown in bold digits. The criterion
one criterion stood out for the Brisbane datasets, which
with the highest misfit (worst correlation) is shown
essentially comprised UCS and BTS data.
faded. As reported by Colmenares and Zoback (2002)
the modified Wiebols-Cook criterion had the best
correlation in three of the five data sets.

TABLE 8.6 The r2 values obtained from curve-fitting the test data from recent tunnelling projects
Rock HB-MD HB HB-3D L W-C C

Hawkesbury 0.776 0.720 0.752 0.780 0.670 0.726


Sandstone 5.63 17.44 15.77 16.39 18.91 21.34

0.101 0.077 0 0.213 0.074 0.08


Brisbane Tuff
36.80 36.55 39.70 37.04 39.90 36.47

Brisbane 0.009 0.002 0.001 0.011 0.352 0.005


Phyllite 18.42 17.33 17.27 17.95 21.87 18.66

Brisbane 0.017 0.011 0.012 0.098 0.350 0.012


Quartzite 35.99 33.57 33.45 38.91 27.14 33.45

Brisbane 0 0 0.105 0 0.454 0


Basalt 28.01 28.01 28.01 29.31 20.69 28.01

Brisbane 0.018 0.018 0.018 0.204 0.013 0.018


Greywacke 41.30 42.95 42.26 51.78 46.97 42.04

No. of times
with smallest
2 0 1 0 2 1
SEE [often
highest r2]

No. of time
with largest
0 0 0 3 2 1
SEE [often
lowest r2]

119
ESTIMATING ROCK MASS PROPERTIES

TABLE 8.7 The misfit values from Colmenares TABLE 8.8 The correlation obtained from
and Zoback (2002) Kwasniewski (2013)
Misfit [MPa] Dataset HB-3D
Dataset
H-B Modified L W-C r2 SEE [MPa]

KTB amphibolite 89.9 91.3 77.8 1 Izumi sandstone 0.9789 6.341


Meta

Dunham dolomite 56.2 27.8 27.4 7 Dunham dolomite 0.9868 5.976

Solenhofen 37.4 23.3 25.5 8 Solenhofen 0.9573 5.422


Sediment

Limestone limestone

Yuubari Shale 13.0 13.7 12.8 9 Yamaguchi marble 0.9900 3.509

Shirahama 8.7 11.9 10.3 10 Mizuho trachyte 0.9773 3.815


Sandstone
11 Manazuru andesite 0.9891 7.532
Number with lowest 1 1 3
misfit 12 Inada granite 0.9961 8.316

Number with highest 2 3 0 13 Orikabe monzonite 0.9880 10.533


misfit
25 Chichibu schist 0.9917 4.145

26 Shirahama 0.9762 3.027


8.3.5 Kwasniewski sandstone
Kwasniewski (2013) collated a comprehensive database 27 Izumi sandstone 0.9498 8.008
of 3D laboratory testing and found that the Hoek-
28 Horonai sandstone 0.9255 2.752
Brown criterion as extended to 3D by Zhang and
Zhu (2007), (HB-3D) was the best fit to 14 of the 31 Yuubari shale 0.9432 2.554
29 datasets. The Mogi linear criterion was the best fit
32 Yamaguchi marble 0.9743 2.580
to six cases and the Mogi power criterion the best fit
to nine cases. The r2 and the SEE for the HB-3D for 33 Westerly granite 0.9622 8.396
each data set is summarised in Table 8.8. Excellent 34 Taga limestone 0.9417 4.145
correlation was found with r2 values ranging from
0.9255 to 0.9976. 35 Locharbriggs 0.9679 1.814
sandstone

8.4 DISCUSSION 36 SSC muddy


siltstone
0.9796 0.636

Several interpretations can be drawn from this


comparison of strength criteria curve-fits to laboratory 37 Westerly granite 0.9906 10.853
test data. 38 KTB amphibolite 0.9900 13.039
• Any of the assessed criteria can provide good fits to 39 Slask sandstone 0.9976 1.328
tight test data, that is test data that does not show
much scatter. 40 Rozbark sandstone 0.9959 1.678
• W hen there is considerable scatter in the test data, 41 Soignies limestone 0.9374 7.949
the correlations were poor across all the criteria.
42 Phra Wihan 0.9948 1.271
• Outlier test data, particularly UTS or BTS values, sandstone
can significantly affect each of the criteria’s
parameters. 43 Phu Phan 0.9963 0.953
sandstone
• W hen tight test data exists for a wide range of
confining stresses, the Hoek-Brown (and its 44 Phu Kadung 0.9952 0.948
variations), the Lade and Wiebols-Cook all provide sandstone
excellent correlations.
45 Phra Wihan 0.9935 0.737
• It is worth noting that curve-fitting a unique set of sandstone
the Lade parameters a, η1 and m was difficult. This
can be appreciated by considering Figure 8.12. Lade 46 Maha Sarakham 0.9712 1.451
rock salt
(1993) showed that small changes in the parameters
give rise to large changes in the strength envelope. 47 SAFOD 0.9775 12.609
granodiorite

120
8 | Intact Rock Test Data

• The Christensen criterion generally gave the poorest • UCS test data alone cannot be used to define any of
fit to the data. It can readily be seen that the the strength envelopes.
Christensen criterion does not particularly represent • The scatter is generally far too great and the range of
test data at high confinements. confining stresses far too limited in the ‘real world’
• The variation in test data swamps the differences laboratory test data from the Sydney and Brisbane
in criteria. tunnelling projects to provide conclusive solutions
• The above observations are not surprising as the to any of the criteria.
criteria share a reasonably similar structure and all • Even when all the criteria provide very good fits to
have elements that are likely to fail at the extremes. the data based on the r2 and SEE values and on the
This is in keeping with Al-Ajmi and Zimmerman overall shape of the envelopes, many of the criteria are
(2005) who also found that the variance is due more poor estimators of strength in the low stress region
to the experimental scatter than to any inherent where many civil structures, such as tunnels, operate.
difficulty in curve fitting.
• The difference between the envelopes in the tensile 8.5 CONCLUSIONS
region is most pronounced in datasets without
tensile strength test results and where the triaxial The question of what is the best fitting strength
data is over a limited range of confining stresses, criterion has been the subject of much research and
is highly variable or is non-existent. These datasets many publications. It will probably continue to be so.
illustrate the limitations of the Hoek-Brown and To paraphrase Christensen (1997) many empirical
Wiebols-Cook criteria. The Hoek-Brown criterion criteria with two or more parameters, such as those
over-predicts the tensile strength and the Wiebols- considered in this paper, would likely fit some sets of
Cook criterion tends to quickly over-predict the test data very well.
triaxial strength. The work carried out confirms this and suggests
• The sharp curvature shown in some datasets is that any of the criteria considered herein could
a result of curve-fitting following the Mostyn- provide reasonable estimates for the intact strength
Douglas method which allows for a variable ‘a’. In of the rocks assessed. However, curve-fitting to derive
this particular case, the curve-fitting results in high values of parameters for strength criteria becomes
‘mi ’ and low ‘a’ values. problematic when there is variability in the test data.
Unfortunately, test data variability is a universal fact.
• Curve fitting can be performed when the test data
comprises only UCS and BTS, though in practical On balance, the Hoek-Brown criterion with
terms any number of alternate envelopes could parameters mi and a derived as recommended by Mostyn
be justified. The Hoek-Brown criterion (and its and Douglas (2000) and extended to 3D where required
variations) appears to produce the more reasonably by following Zhang and Zhu (2007) produces the better
shaped envelopes. fit to the majority of the rock types considered.

1.E+06
Mostyn & Douglas (2000) 1.E+06
Artificial Data Hawkesbury Sandstone
1.E+05
η1 = 1.396E+04 1.E+05
η1 = 1.10E+04
1.E+04
m= 0.94 1.E+04 m= 0.500
I13 / I3 - 27

I13 / I3 - 27

1.E+03
1.E+03

1.E+02
1.E+02
UTS BTS
1.E+01 UCS UCS
1.E+01
Triaxial Triaxial
1.E+00
1.E+00
1.E-04 1.E-03 1.E-02 1.E-01 1.E+00
1.E-04 1.E-03 1.E-02 1.E-01 1.E+00
pa / I1
pa / I1

FIGURE 8.12 Examples of the data used to estimate the Lade parameters, m and η1, for the (left) artificial data set (mean
values only used) and for the (right) full laboratory test results for Hawkesbury Sandstone.

121
ESTIMATING ROCK MASS PROPERTIES

122
9 | GSI Charts

9 GSI CHARTS

9.1 OUTLINE Marinos (2010) presents six GSI charts for flysch,
This chapter continues from Chapter 4 by looking molasse, limestone, ophiolite and disturbed / weathered
in detail at the Geological Strength Index (GSI) gneiss based on data collected during the construction
classification charts. of 62 tunnels along the Egnatia Highway in Northern
Greece. For completeness, these charts are reproduced
The GSI is intended to cover a wide range of geological here in Figure 9.3 to Figure 9.8.
conditions, including weak, foliated and sheared rock
masses (Hoek et al. 1998) and heterogeneous rock The latest chart is that published by Hoek et al.
masses, such as flysch (Marinos and Hoek 2001). These (2013) for jointed, blocky rock mass for tunnels of
two charts are shown in Figure 9.1 and Figure 9.2. about 10 m span and slopes < 20 m high.

FIGURE 9.1 GSI chart for general rock mass (left) and extended to cover schist (right) (Hoek et al. 1998)

123
ESTIMATING ROCK MASS PROPERTIES

FIGURE 9.2 GSI chart for flysch from Marinos and Hoek (2001)

FIGURE 9.3 GSI chart for flysch; from Marinos (2014)

124
9 | GSI Charts

FIGURE 9.4 GSI chart for molasse at depth; from Hoek et al. (2005)

FIGURE 9.5 GSI chart for molasse at surface; from Marinos (2010)

125
ESTIMATING ROCK MASS PROPERTIES

FIGURE 9.6 GSI chart for ophiolites; from Marinos (2010)

126
9 | GSI Charts

FIGURE 9.7 GSI chart for limestone; from Marinos (2010)

FIGURE 9.8 GSI chart for gneiss; from Marinos (2010)

127
ESTIMATING ROCK MASS PROPERTIES

Apart from the curved contours for gneiss (Figure to bottom-right as the rock mass deteriorates. Marinos
9.8), the charts are essentially the same. They effectively et al. (2006) and Marinos (2010) show the same
describe specific geological conditions as they apply diagonal trend.
to each of the rock masses in more detail than the Upon reflection, the diagonal trend should not
descriptions given in the generic GSI chart. be surprising. Often the blockiness of the rock mass
Carter and Marinos (2014) prepared a matrix (vertical axis) is related to the surface quality of a rock
showing the relationships between the GSI charts and mass (horizontal axis). Hence, the GSI chart’s axes are
the Hoek-Brown strength criterion parameters of σc not necessarily independent. Marinos and Hoek (2000)
and mi for different lithology. The matrix highlights also observed that not all combinations are possible and
the importance of first developing the geological model exclude the top right and bottom left cells in the GSI
before estimating the GSI. chart as can be seen in Figure 9.3 to Figure 9.9.
An attempt to collate the data back into the generic It is also interesting to observe the restricted range of
GSI chart is shown in Figure 9.9. The trend in GSI GSI when the limitations suggested by Brown (2008);
is obvious: it trends down the diagonal from top-left Carter et al. (2007); (2008); Carvalho et al. (2007);

FIGURE 9.9 GSI chart for the five rock masses presented in Marinos (2010).

128
9 | GSI Charts

Diederichs (2007) are superimposed. That is, the GSI grained, quartzose sandstone which had been deposited
used to predict the Hoek-Brown strength parameters in 0.5 to 5 m thick beds, typically 1 to 2 m, that exhibit
should be limited to approximately the range 30 – 65 either massive or cross-bedded facies. Paraphrasing the
(see Section 5.6.2). This is overlaid onto the diagonal words of McNally (2000), Hawkesbury Sandstone
trend in Figure 9.9 as the pale purple shaded region. is singularly well-suited to excavations because of a
To further explore if this trend in GSI is generally number of geological characteristics:
applicable, four very different rock masses are assessed. • It is weak enough to be easily (machine) excavated
yet strong enough to require minimal support
• The Hawkesbury Sandstone and Ashfield Shale of
Sydney, Australia • It is horizontal and thickly bedded and relatively
undeformed
• The Greenland Group greywacke and argillite of
Reefton, South Island, New Zealand • Jointing is wider spaced than bedding planes, sub-
vertical and orthogonal.
• Otago Schist of Macraes, South Island,
New Zealand
9.2.2 Ashfield Shale
The quantification of GSI using the RMR joint
The Ashfield Shale has a maximum thickness of
condition, JCond89 , and RQD as proposed by Hoek et
approximately 60 m and grades from a claystone,
al. (2013) is also explored.
siltstone to a siltstone / fine grained sandstone laminate
(MacGregor 1985). It has been extensively worked
9.2 SYDNEY throughout Sydney as a source of brick-making clay
Borehole data collected for several tunnelling projects (Branagan and Packham 2000).
in Sydney has been used to characterise Sydney’s The Ashfield Shale comprises well developed thinly
Hawkesbury Sandstone and Ashfield Shale based on laminated rock with wider spaced bedding from about
the Sydney classification system (Pells et al. 1998). The 0.1 - 0.5 m to 2 m vertical spacing. The bedding
database includes information from the Ocean Outfalls, planes are sub-horizontal dipping typically 0 to 5° and
Sydney Harbour Tunnel, M2, Eastern Distributor, M5 persistent over tens to hundreds of metres.
East, Cross City, cable tunnels, Epping to Chatswood
rail link, Lane Cove, CBD Metro (project was not 9.2.3 Sydney rock mass classification
built), Wynyard Walk and the Northwest Rail link The classification system for Sydney sandstone and
projects. Detailed borehole logging from the last three shale, Pells et al. (1998) which updated Pells et al. (1978),
projects has been especially used. was intended to assist in the design of foundations on
rock in the Sydney area. The classification system is
9.2.1 Hawkesbury Sandstone based on rock strength, defect spacing and allowable
Central Sydney is founded on the Hawkesbury seams as shown in Table 9.1. All three factors must
Sandstone, which varies in thickness from 30 m in the be satisfied. Seams include clay, fragmented or highly
lower Blue Mountains to 240 m in the Hawkesbury weathered zones.
River district (Branagan 2000). The Hawkesbury Pells et al. (1998) recommended that the zone of
Sandstone is often described as a medium to coarse rock being classified be “over a length of core of similar

TABLE 9.1 Sydney Classification System (Pells et al. 1998)


CLASS UCS DEFECT SPACING (a)
ALLOWABLE SEAMS
[MPa] %
[mm] Description
I > 24 > 600 Widely spaced < 1.5
Sandstone

II > 12 > 600 Widely spaced <3


III >7 > 200 Moderately spaced <5
IV >2 > 60 Closely spaced < 10
V >1 NA NA
I > 16 > 600 Widely spaced <2
II >7 > 200 Moderately spaced <4
Shale

III >2 > 60 Closely spaced <8


IV >1 > 20 Very closely spaced < 25
V >1 NA NA
Defect spacing based on ISO/DIS 14689 & ISRM suggested methods replaced the degree of fracturing terms in the Pells et al.
(a) 

(1978) paper

129
ESTIMATING ROCK MASS PROPERTIES

characteristics”. That is, “the classification system be


applied to portions or units of rock mass having similar
UCS, defect spacing and seam characteristics” (Bertuzzi
and Pells 2002). An example profile of a mapped face
was provided in that paper to clarify the intended
method of applying the classification system. This is
reproduced in Figure 9.10.

FIGURE 9.10 The wrong and right way to classify (Bertuzzi and Pells 2002)

While the Sydney Classification System was not classification was updated in 1998. However, the defect
originally intended for tunnelling, it does represent a spacings appropriate for foundation problems have
good method for communicating rock mass quality been found to cover a too narrow range for tunnelling,
in Sydney sandstone and siltstone. It is also useful for which needs to consider spacing in three dimensions.
linking values put forward for designs with measured Typical minimum spacing for defects, bedding and
and back-figured parameters from existing excavations, jointing is therefore suggested in Table 9.2. It is not
e.g. Bertuzzi and Pells (2002); Clarke and Pells (2004). intended that this replaces the Sydney Classification
Hence, it is often used as the basis for tunnelling System but rather used as a guide as to what to expect
projects in Sydney and was being so used when the in the tunnelling environment.

TABLE 9.2 Suggested conditions for tunnelling projects


CLASS UCS DEFECT SPACING (m) ALLOWABLE SEAMS
[MPa] DEFECTS TYPICAL %
Bedding Joints

I > 24 > 0.6 > 1.5 >2 < 1.5


Sandstone

II > 12 > 0.6 >1 >1 <3


III >7 > 0.2 > 0.5 > 0.5 <5
IV >2 > 0.06 > 0.2 > 0.2 < 10
V >1 NA NA
I > 16 > 0.6 >1 >1 <2
II >7 > 0.2 > 0.5 > 0.5 <4
Shale

III >2 > 0.06 > 0.1 > 0.2 <8


IV >1 > 0.02 > 0.1 > 0.1 < 25
V >1 NA NA

130
9 | GSI Charts

9.2.4 Field data validation • Field estimated strength terms are those of AS 1729-
Recent tunnelling projects in Sydney have provided a 1993 Geotechnical site investigations
large database of drill core logs, comprising 6 km of logs • Defect spacing based on ISO/DIS 14689 & ISRM
containing over 7200 records of defect characteristics suggested methods
as well as field estimated strengths, point load index
tests and laboratory UCS tests. The data has been
• Defect characteristics follow AS 1729-1993
Geotechnical site investigations
collected by several different geotechnical organisations
and various geotechnical engineers and engineering This data validates the bedding and joint spacing for
geologists. Hence, any bias of particular geotechnical tunnelling projects suggested in Table 9.2.
organisations or individuals is likely to be balanced out In terms of orientation, the recorded data shows
in the collated data. bedding is obviously sub-horizontal (to 10°) in both
The database has been accessed to assess the the sandstone and shale whereas joint patterns differ
variability in intact strength and the spacing, (Figure 9.11). In the sandstone, two sub-vertical
characteristics and orientation of defects for each joint sets are apparent - striking north-northeast and
particular class of sandstone and shale. The variability east-southeast. While these two orientations are also
is presented in Figure 9.13 and Figure 9.14 and in apparent in the shale, other joint sets do occur dipping
Table 9.3 and Table 9.4. The terms used in these tables between 30 to 50° in various directions.
are from the following standards and guidelines.

FIGURE 9.11 Equal area projections for Hawkesbury Sandstone on the left (4675 defects) and Ashfield Shale on the right
(1338 defects).

FIGURE 9.12 Subset of the data in Figure 9.11 just showing joints for Hawkesbury Sandstone on the left and Ashfield Shale
on the right.

131
ESTIMATING ROCK MASS PROPERTIES

FIGURE 9.13 Distribution of logged spacing and aperture for bedding planes and joints for Hawkesbury Sandstone

FIGURE 9.14 Distribution of logged spacing and aperture for bedding planes and joints for Ashfield Shale

132
9 | GSI Charts

TABLE 9.3 Variability in strength and spacing as recorded in the drill hole database
STRENGTH TYPICAL SPACING TOTAL CORE
RANGE [m] SAMPLE
CLASS AXIAL Is(50) [MPa] LENGTH [m]
Field estimate Tests Range Ave Bedding Joint
I High 1575 1.0 – 3.0 1.7 0.6 - > 6.0 0.6 - > 6.0 2478
Sandstone

II High 827 0.3 – 3.0 1.5 0.6 - 6.0 0.2 - 6.0 890
III Medium to High 625 0.3 – 3.0 1.4 0.2 - 2.0 0.2 - 2.0 590
IV Low to High 260 0.1 – 3.0 1.0 0.06 - 2.0 0.06 - 2.0 241
V Very low to High 41 <0.03 – 3.0 0.9 0.06 - 2.0 0.06 - 2.0 42
I Medium to High 69 0.3 – 3.0 1.5 0.2 - 6.0 0.2 - 6.0 471
II Low to High 130 0.1 – 3.0 1.3 0.2 - 6.0 0.2 - 6.0 499
Shale

III Low to High 100 0.03 – 3.0 1.0 0.06 - 2.0 0.06 - 2.0 271
IV Very low to High 31 <0.03 – 3.0 0.5 0.06 - 2.0 0.06 - 2.0 130
V Very low to High 10 <0.03 – 3.0 0.08 0.06 - 2.0 0.06 - 2.0 96

TABLE 9.4 Variability in the defect characteristics as recorded in the drill hole database
CLASS DEFECT DEFECT CHARACTERISTICS DEFECTS
TYPE Rough Shape Aperture (mm) Infill RECORDED

I BG Rough Planar – Undulating Clean to 1 - 5 Clay 1066


JN Rough Planar – Undulating Clean - 315
II BG Rough Planar - Undulating Clean to 1 - 10 Clay 952
JN Rough Planar Clean to Veneer Clay - Fe 230
Sandstone

III BG Rough Planar Clean to 1 - 50 Clay - Fe 1369


JN Rough Planar Clean to Veneer Clay - Fe 511
IV BG Rough Planar 1 – 50 Clay - Fe 897
JN Smooth - Planar Clean to 1 - 5 Clay - Fe 376
Rough
V BG Rough Planar - Undulating 1 – 100 Clay - Fe 117
JN Smooth - Planar - Undulating 1–5 Clay – Fe 73
Rough
I BG Smooth - Planar Clean to Veneer Clay 106
Rough
JN Smooth - Planar Clean - 217
Rough
II BG Smooth - Planar Clean to 1 - 5 Clay - Fe 314
Rough
JN Smooth - Planar Clean to Veneer Clay - Fe 467
Rough
Shale

III BG Smooth Planar Clean to 1 - 50 Clay - Fe 420


JN Smooth - Planar - Undulating Clean to Veneer Clay - Fe 411
Rough
IV BG Smooth Planar 1 – 50 Clay - Fe 205
JN Smooth - Planar - Undulating Clean to 1 – 5 Clay - Fe 108
Rough
V BG Smooth Planar 1 – 100 Clay - Fe 124
JN Smooth - Planar - Undulating Clean to 1 – 5 Clay - Fe 18
Rough

133
ESTIMATING ROCK MASS PROPERTIES

The typically observed defect characteristics Persistence of cross-bedding is controlled by


and those which can occur but represent adverse sandstone bed thickness. Geotechnical models should
conditions are summarised in Table 9.5 and Table consider and include adverse conditions, particularly
9.6 for Hawkesbury Sandstone and Ashfield Shale, that of defect persistence.
respectively. Spacings are given in Table 9.2.

TABLE 9.5 Typical and adverse conditions of defects in Hawkesbury Sandstone


PARAMETER TYPICAL ADVERSE
Persistence [m] 1. > 20 2. > 20
Roughness Rough Slightly rough
Bedding Shape Undulating Planar
Aperture [mm] <1 > 10
Infill None / limonite Sandy clay
Persistence [m] 3. > 10 4. > 10
Class I to V

Roughness Rough Slightly rough


Cross-bedding Shape Undulating Planar
Aperture [mm] <1 <1
Infill None / limonite None
Persistence [m] 5.5 6. >10
Roughness Rough Slightly rough
Jointing Shape Planar Planar
Aperture [mm] <1 > 10
Infill None / limonite Sandy clay

TABLE 9.6 Typical and adverse conditions of defects in Ashfield Shale


PARAMETER TYPICAL ADVERSE
Persistence [m] 7. > 20 8. > 20
Roughness Slightly rough Smooth
Bedding Shape Undulating Planar
Class I, II & III

Aperture [mm] <1 > 10


Infill None / limonite Silty clay
Persistence [m] 9. 3 10. > 5
Roughness Slightly rough Smooth
Jointing Shape Planar Planar
Aperture [mm] <1 > 10
Infill None / limonite Silty clay
Persistence [m] 11. >20 12. >20
Roughness Slightly rough Smooth
Bedding Shape Undulating Planar
Class IV & V

Aperture [mm] <1 50


Infill Clay coating Silty clay
Persistence [m] 13. 3 14. >5
Roughness Slightly rough Smooth
Jointing Shape Planar Planar
Aperture [mm] <1 50
Infill Clay coating Silty clay

134
9 | GSI Charts

9.2.5 Classification presented in McQueen (2004), the relationship, σH =


In order to calculate the Q value an estimate of the in (1.5 to 2.5) + 2.0σV, is proposed as shown in Figure 9.15.
situ stress is required, or more correctly, the ratio of This means that the typical in situ principal stresses for
intact rock strength to stress. Enever (1999) collated depths less than 50 m (stresses at specific locations may
in situ stress data that suggested the following stepped be different) can be assumed to be:
profile for the upper bound major horizontal stress (σH): • σV up to 1.2 MPa
• Approximately 2.5 MPa above vertical stress (σV) • σH ≈ (1.5 to 2.5) + 2.0σV MPa, i.e. up to 5 MPa.
for depths less than 20 m It should be noted that while the high horizontal
• Approximately 6.5 MPa above σV for 20 – 200 m stress ratio can be sustained in the better quality rock
depths masses of say Sandstone Classes I, II & III and Shale
• Approximately 15 MPa above σV for 200 – 1200 m Classes I and II, they cannot be sustained in the poorer
depths quality rock masses.
Rather than adopting this profile with seemingly Values of GSI and Q for the five classes of
arbitrary steps, Pells (2004) suggested σH can be related Hawkesbury Sandstone and Ashfield Shale are
to σV to simplify the design process as: σH = 1.5 + (1.2 summarised in Figure 9.16 and Figure 9.17, respectively.
to 2.0)σV. Adopting this same approach to the data

FIGURE 9.15 Major horizontal stress versus depth with the design line σH = 1.5 + 2.0σV and an upper bound σH = 2.5 + 2.3σV
shown; after McQueen (2004)

135
ESTIMATING ROCK MASS PROPERTIES

FIGURE 9.16 GSI values for Hawkesbury Sandstone rock mass; the arrow shows the progression from Class I to Class V

FIGURE 9.17 GSI values for Ashfield Shale rock mass; the arrow shows the progression from Class I to Class V

136
9 | GSI Charts

9.3 REEFTON • Bedding is typically planar, smooth and clean;


Bedding in greywacke φ’ = 30° and c’= zero; Bedding
9.3.1 Geology in argillite φ’ = 24° and c’= zero.
The Reefton gold deposit is in the northwest of the • Crush zones, shears and faults are typically planar
South Island of New Zealand, approximately 7 to undulating, smooth to rough and infilled with
km southeast of the township of Reefton (Figure clay and/or rock fragments with widths of typically
9.18). The area is a highly active seismic zone, being 5 to 200mm; φ’ = 20-25° and c’= zero.
approximately 35 km from the Alpine Fault. The • Joints are typically planar, smooth to rough,
Reefton gold deposit has been mined intermittently typically clean although some surfaces display
since the late 1870s. OceanaGold open pit mined the limonite veneers; φ’ = 33° and c’= zero.
Reefton deposit between 2002 and 2016.
The Reefton mine area comprises Ordovician-aged 9.3.2 Field data
interbedded high strength greywacke and argillite (9) of The rock mass conditions at Reefton have been
the Greenland Group. This sequence is folded into a assessed using available core photographs and logging
series of tight, north-south trending, gently plunging data from 31 geotechnically logged boreholes.
folds. The shape of the folds results in the overall Each borehole was assessed using the core
bedding generally dipping between 20 and 70° towards photographs and logs. Borehole intervals assessed to
the west, although the local orientation is dependent have similar geotechnical characteristics were then
upon the folding. The folds are offset by several faults. interrogated to provide histograms of lithology, RQD,
The following summary of the defect characteristics is weathering, alteration, field estimated strength and
based on mapping. defect characteristics.

FIGURE 9.18 Map of Reefton Goldfield showing location of Globe-Progress Open Pit Gold Mine and other historically
important mining centres (modified from OceanaGold Ltd, 2008)

(9)
 verage uniaxial compressive strength (UCS) of greywacke is 80 MPa (varying between 20 and 318 MPa) and for argillite, 50 MPa measured
A
perpendicular to bedding and 15 MPa measured parallel to bedding.

137
ESTIMATING ROCK MASS PROPERTIES

Following this process, five classes of rock mass


were defined (A to E) with each class showing similar
geotechnical characteristics (Table 9.7).

TABLE 9.7 Reefton rock mass classes


CLASS GENERAL DESCRIPTION INTACT ROCK STRENGTH RQD

A Lithified rock with widely spaced Medium to high 90 – 100%


defects and rare shearing

B Fractured rock with frequent defects Low to medium 60 – 70%


and rare shearing

C Highly fractured rock with frequent Low to medium 40 – 60%, with zones of 0 –
defects and some shearing 10%

D Fractured to fragmented rock with Extremely low to medium 30 – 40% with frequent zones
frequent shearing of low quality 0 – 10%

E Fragmented and sheared low grade Extremely low – low rock <15%
rock – soil

9.3.3 Classification
Values of GSI for the five classes of rock types found at
Reefton, which comprise greywacke and argillite, are
summarised in Figure 9.19.

FIGURE 9.19 GSI values for Reefton greywacke and argillite rock mass; the arrow shows the progression from Class A to Class E

138
9 | GSI Charts

9.4 MACRAES The schist comprises a sequence of interlayered and


gradational psammitic and pelitic lithologies that have
9.4.1 Geology been derived by metamorphism of Mesozoic sandstone
The Macraes Flat goldfield is located approximately and mudstone respectively. The rocks are strongly
60km north of Dunedin in the Otago region of foliated and depending on origin are either light grey,
the South Island of New Zealand (Figure 9.20). quartz rich and laminated (psammite) or dark grey to
OceanaGold has been open pit mining at Macraes Flat, green, micaceous and finely laminated (pelite).
Otago, New Zealand since October 1990 in a series The intact rock strength does vary across the rock
of pits along strike of the Hyde-Macraes Shear Zone. types but it is not distinctive and is very dependent
The Shear Zone is part of the extensively deformed upon the degree of weathering and whether it is
Otago-Haast Schist Belt. The structural geology of tested parallel or perpendicular to foliation. Typically,
the area is dominated by two main orthogonal fault the schist is of high strength (UCS = 25 to 50 MPa)
sets, striking to the north and east (10). The Shear Zone perpendicular to foliation but is of medium strength
dips at about 15 to 20° to the east and is approximately parallel to foliation (UCS = 10 to 25 MPa).
100m thick. It is formed within deformed pelitic schist
so that its continuous surface-boundary faults also 9.4.2 Field data
approximate lithological boundaries. The geotechnical model of the Macraes area is
Tectonic displacement of the Shear Zone has been summarised below.
inferred to be hundreds of metres. The strain associated • Foliation, foliation shears and faults showing a
with this displacement was probably concentrated broad range of dips from flat to moderately dipping
within the intra-shear pelite which could absorb strain towards the east (average 15 to 45°/100°) – see the
more readily than the coarsely grained psammite. left stereoplot in Diagram 2. The set parallels the
Hyde Macraes Shear Zone.

FIGURE 9.20 Map of Macraes showing location of Hyde-Macraes Shear Zone

(10)
The directions quoted are relative to Macraes’ mine grid, which is 45° west of true north and approximately 67½° west of magnetic north.

139
ESTIMATING ROCK MASS PROPERTIES

• Mine and regional scale faults that moderately to • Jointsand faults that are moderately to steeply
steeply dip towards the east (≈60°/090°) – Set 1 on dipping towards the north and south (≈ 55-75°/025°)
the right stereoplot in Figure 9.21. These faults are – Set 2 on the right stereoplot in Figure 9.21.
often infilled with clay or breccia to 100 mm thick. The same process as that for Reefton (Section 9.3.2)
• Batter and mine scale faults and shears showing a was followed to define four classes of schist rock mass
broad range of dips from flat to moderately dipping (A to D) with each class showing similar geotechnical
towards the west (average ≈ 50°) – Set 3 on the right characteristics (Table 9.8).
stereoplot in Figure 9.21. This fault set is typically
truncated by the easterly dipping faults, described
in the second dot point above.

FIGURE 9.21 Stereoplots showing typical orientations of foliation (left) and faults / shears and joints (right) at Macraes

TABLE 9.8 Macraes rock mass classes


CLASS GENERAL DESCRIPTION INTACT ROCK STRENGTH RQD

A Lithified rock with frequent defects and High 75 – 100%


rare shearing

B Fractured rock with frequent defects Low to high 60 – 80%


and some hearing

C Fractured to fragmented rock with Low to high 40 – 60%, with zones of


frequent shearing 0 – 10%

D Fractured / sheared rock Extremely low to very low 10 – 20%

140
9 | GSI Charts

9.4.3 Classification
Values of GSI for the four classes of Otago Schist at
Macraes are summarised in Figure 9.22.

FIGURE 9.22 GSI values for Otago Schist rock mass at Macraes; the arrow shows the progression from Class A to Class D

9.5 QUANTIFICATION OF GSI A quantified GSI was calculated following the


Hoek et al. (2013) propose quantifying GSI with RQD methodology of Hoek et al. (2013) for the four rock
and JCond89. A similar quantification was proposed masses described in the preceding sections. The data
by Cai et al. (2004). This was explored in Chapter 4. was used to calculate the JCond89 rating which was
Here, the proposed quantification method is tested found to vary for each rock mass class as shown in
against the four rock masses. Figure 9.23, for example.
The GSI values based on:
Defect Condition • the chart assessed by visual interpretation of the
50
Sandstone II
100%
rock mass at a scale suitable for a tunnel or multiple
45 90%
40 80%
benches in an open pit mine using the sketches on
35 70%
the left-hand side of the GSI charts shown in Figure
30 60% 9.16 to Figure 9.22
25 50%
• the quantification approach of Hoek et al. (2013)
%

20 40%
15 30%
are listed in Table 9.9 and graphed in Figure 9.24.
10 20% The overall correlation between the two sets of values
5 10% is fair as measured by r2 = 0.68. As can be seen, the
0
1 3 5 7 9 11 13 15 17 19 21 23 25 27 29
0% greater the GSI the poorer the correlation. By way
of comparison, the example provided in Hoek et al.
Rating
(2013) yielded r2 =0.95.
FIGURE 9.23 Variation in JCond as calculated for Most of the data from the two methods plots within
Hawkesbury Sandstone Class II a band defined around GSI ± 10. This is in keeping

141
ESTIMATING ROCK MASS PROPERTIES

the data plotted by Russo (2009) as shown in Figure suggests that the correlation between the two methods
9.25. A variation of GSI ± 10 is reasonable considering to estimate GSI becomes poorer as RQD increases.
the natural variability of rock masses. There does not appear to be a clear trend with JCond89.
The noticeable exception is the poorer correlation Priest and Hudson (1976) showed that RQD is
for the Ashfield Shale and for the top class of both the insensitive when the mean discontinuity spacing is
Greenland Group greywacke and Otago Schist. The more than 0.3 m. This is in keeping with the findings
Ashfield Shale is a laminated to thinly bedded siltstone reported by Palmström (2005) that RQD is insensitive
/ laminite but its RQD from borehole core of Class I and to changes in joints per cubic metre (Jv) greater than
II is excellent. Both the Greenland Group greywacke 5/m. The RQD is therefore insensitive to the range of
and Otago Schist are foliated metamorphic rocks with lamination and bedding spacing of these three rock
the better quality rocks having high RQD. High RQD masses. An experienced practitioner using the GSI
tends to give the poorer correlation as seen in Figure chart would be able to identify this issue and choose
9.26 which presents the same data as in Figure 9.25 but an appropriate value.
grouped according to RQD and JCond89. The figure

TABLE 9.9 Comparison of quantified GSI and that obtained from its chart

CLASS TYPICAL RQD JCond89 CHART GSI QUANTIFIED GSI


(Mean, Range) (Mean, Range) (Mean, Range)

I 90 – 100 20 (11 – 33) 75 (65 – 85) 70 (57 – 75)

II 90 – 100 15 (9 – 22) 65 (55 – 75) 63 (55 – 73)


Hawkesbury
Sandstone

III 75 – 100 12 (8 – 19) 55 (45 – 65) 58 (55 - 68)

IV 50 – 75 7 (4 – 16) 45 (35 – 55) 42 (34 – 56)

V < 50 6 (3 – 12) 35 (25 – 45) 20 (10 – 38)

I 90 – 100 22 (12 – 26) 55 (45 – 65) 73 (60 - 80)


Ashfield Shale

II 80 – 100 22 (7 – 26) 50 (40 – 60) 73 (55 - 78)

III 40 – 70 18 (7 – 23) 40 (30 – 50) 55 (36 – 65)

IV < 40 10 (2 – 21) 30 (25 – 40) 25 (12 – 45)

V < 40 6 (1 – 19) 25 (15 – 30) 16 (8 – 35)

A 60 – 100 17 (13 – 20) 80 (75 – 85) 57 (47 – 67)


Greenland Group

B 30 – 70 17 (9 – 20) 55 (45 – 65) 43 (34 – 55)


Greywacke /
argillite

C 20 – 60 15 (9 – 19) 40 (35 – 45) 42 (30 – 55)

D 10 – 40 11 (7 – 19) 30 (25 – 35) 23 (16 – 35)

E < 15 13 (1 – 21) 20 (15 – 25) 16 (5 – 24)

A 60 – 100 18 (7 – 25) 75 (70 – 80) 66 (50 – 80)


Otago Schist

B 30 – 70 18 (6 – 22) 60 (45 – 70) 53 (32 – 62)

C < 40 11 (2 – 22) 30 (25 – 35) 27 (11 – 46)

D < 20 12 (4 – 19) 15 (12 – 20) 23 (10 – 35)

142
9 | GSI Charts

100

Quantified GSI = Chart GSI ± 10

80 r2=0.679
r2

60
Quantified GSI

40

Greenland Group Greywacke / argillite


20 Otago Schist
Ashfield Shale
Hawkesbury Sandstone

0
0 20 40 60 80 100
Chart GSI

FIGURE 9.24 Comparison of quantified GSI and the GSI assessed from the chart. The highlighted region shows a band
Quantified GSI = Chart GSI ± 10.

FIGURE 9.25 Comparison of a quantified and chart GSI from Russo (2009). According to Russo (2009), the red triangles
are data taken by him from papers by Hoek; and the black circles are data from his mapping. Most points fall within a band
defined by GSI ± 10.

143
ESTIMATING ROCK MASS PROPERTIES

FIGURE 9.26 Comparison of quantified and chart GSI. This is the same data as Figure 9.25 except grouped according to
RQD (left) and JCond89 (right).

To see whether block volume would offer a better values of the calculated Jc and Vb are listed in Table
indication of blockiness, the method adopted by Cai 9.10 for Classes I to IV. There was insufficient data for
et al. (2004) was followed to calculate the GSI for the Class V materials.
Hawkesbury Sandstone and Ashfield Shale found in The results suggest that the method of Cai et al.
Sydney. The blocks in the Hawkesbury Sandstone and (2004) to quantify GSI produces results consistent
Ashfield Shale are typically formed by the two pairs with those assessed using the chart for Hawkesbury
of sub-vertical joints intersecting the sub-horizontal Sandstone and Ashfield Shale. It is thought that this
bedding partings. Cross-beds are occasionally involved. is largely a function of Vb being a better measure of
Distributions of spacings for bedding plane partings blockiness than RQD.
and joints are shown in Figure 9.27. The typical

FIGURE 9.27 Distribution of logged spacing for bedding planes and joints for Hawkesbury Sandstone and Ashfield Shale.

144
9 | GSI Charts

TABLE 9.10 Comparison of the GSI quantified following the method of Cai et al. (2004) and that
estimated using the chart.
CLASS HAWKESBURY SANDSTONE ASHFIELD SHALE

Jc Vb GSI Jc Vb GSI
[m3] [m3]
Cai et. al Chart Cai et. al Chart

I 2.71 8.7 80 65 – 85 2.31 3.8 75 45 – 65

II 1.68 3.0 75 55 – 75 0.89 0.7 55 40 – 60

III 1.02 1.4 60 45 – 65 0.56 0.2 45 30 – 50

IV 0.63 0.3 50 35 - 55 0.26 0.1 40 25 - 35

9.6 CONCLUSIONS
The GSI classification of different rock masses exhibits It is therefore recommended that the quantified
the same trend. Namely, the GSI value follows a GSI approach of Hoek et al. (2013) be used to
diagonal line from top-left to bottom-right as the rock supplement and check the visually assessed chart GSI.
mass deteriorates. Further, it is evident that the range A difference between the two approaches of GSI ± 10 is
of typical GSI values is quite restricted. probably likely.
The correlation between the GSI assessed from The two approaches are needed because of the
its chart and the quantified GSI following the complexity of geology and the natural variation of
methodology of Cai et al. (2004) and that of Hoek et al. rock masses. Consider whether using a quantified
(2013) was found to be fair for the combined datasets GSI approach directly from borehole data without
from four rock masses detailed in this chapter. Most of developing a geotechnical model is valid. Statistical
the data plots within ± 10 from the two methods. analysis could be performed on the quantified GSI
The rock mass that has poorer correlation between values which will give an unfounded impression of
the quantified and chart GSI, also has a mismatch accuracy. Yet the purely visual GSI chart approach
between RQD and the problem scale. The ability of is subjective and dependent on the practitioner’s
the quantified approach to estimate GSI reduces as experience. It is very difficult to visualise a rock mass
RQD, or for that matter Jv or Vb as noted by Hoek et solely from borehole logs.
al. (2013), becomes less able to measure the blockiness As Hoek in the Second Glossop Lecture
of a rock mass. elegantly wrote: “A good engineering geologist and a
It might be possible to address this shortcoming good geotechnical engineer, working as a team, can
of the quantified GSI approach by adopting the usually make realistic educated guesses for each of the
suggestion of Cundall et al. (2008) and express the parameters required for a particular engineering analysis”
vertical axis “as the number of blocks across the scale (Hoek 1999).
of interest” rather than RQD/2. This means that the It is also suggested that different GSI values may
problem scale becomes an explicit input parameter be needed to cater for different analysis, particularly
to GSI. However, this does not address the issue that numerical methods.
block size alone does not capture the degree to which
the rock blocks are interlocked, which is an important
part of the descriptions in the GSI chart.

145
ESTIMATING ROCK MASS PROPERTIES

146
TUNNELS IN
10 | Tunnels in Hawkesbury Sandstone Case Studies

10 HAWKESBURY
SANDSTONE
CASE STUDIES

10.1 OUTLINE size. Mesh elements were graded so that immediately


This chapter presents nine case studies where stress surrounding the excavation they are less than 0.25 m
induced failure has been inferred from observations of and typically less than 0.1 m in size.
tunnels excavated within Hawkesbury Sandstone. The • In situ stress was not specifically measured at any
case studies, which are listed below, cover a range of of the specific locations where failure occurred.
conditions: Sensitivity analyses testing in situ stress values
could be undertaken to assess the variability in
• Construction dates from the 1880s to 2006 model predictions of the principal stresses that were
• Spans from 3.2 to 12.5 m present at the time of the failures. However, in these
• Depths from 17 to 124 m elastic models changes in the assumed in situ stress
• Drill & blast, roadheader and TBM excavation. would simply be reflected in the predicted principal
The aim of the work presented in this chapter is to stresses: the points on a principal stress graph (σ1 v
use numerical modelling to assess the likely magnitudes σ3) would move depending on the assumed in situ
of the principal stresses that were present at the time stresses. While this may result in a better- looking
of the failures and to compare these to the rock mass result, without specific in situ stress measurements,
strength criteria. it would largely be a matching exercise. Hence, the
As each of the case studies can be effectively viewed base case in situ stress was modelled as that typical
as a plane cross-section that remains constant over of Sydney to be consistent with the adopted mass
a substantial length, two-dimensional, plane strain modulus and the testing database as shown in
analysis is suitable. Further, as the intention is to Figure 10.1 and in Bertuzzi (2014); Bertuzzi (2014);
assess the applicability of rock mass strength criteria, Pells (2004).
there is no requirement to explicitly include many „„ σv = γ multiplied by depth, where γ = 24 kN/m
3

geological structures into the models. Hence, the two- „„ σH = σNS = 1.5 + 2σv (as per Chapter 9.2.5)
dimensional finite element analysis program Phase2 (11) „„ σh = σEW = 0.7σH (as shown in Figure 10.1)
is used.
• urther, in keeping with the typical direction of
F

Each of the linear elastic, finite element models principal in situ stresses in and around Sydney, the
comprised: major horizontal stress, σH, is modelled to act in
• Hawkesbury Sandstone as an elastic, isotropic the northeast to southwest orientation unless noted
material with E mass = 2000 MPa and ν = 0.2. These otherwise (McQueen 2004; Pells 2004).
are typical values for this rock mass as presented in The finite element models are interrogated for the
Chapter 6. values of major and minor principal stresses (σ1 and σ3)
• The influence of localised bedding planes was in and around the failure zone(s) (12). The values of σ1
assessed by running models that included them as and σ3 within the failure zone are noted as failed points
and where mapped. The bedding plane stiffness is while the values outside the failure zone are noted as
modelled as k n = 100 GPa/m and k s = 10 GPa/m. stable points. These failed and stable points are then
These are typical values of stiffness as presented in compared to proposed rock mass strength envelopes of
Chapter 3. Hawkesbury Sandstone.
• The boundaries of the finite element mesh extend a
distance of at least three times the excavated tunnel

(11)
RocScience Version 8.01
(12) 
As Hawkesbury Sandstone is not a particularly high strength rock (σc is typically 20 to 40 MPa), based on the discussion presented in Chapter 2, Section
2.4.2 the rock mass strength is considered to be independent of σ2.
147
ESTIMATING ROCK MASS PROPERTIES

FIGURE 10.1 Relationship of major horizontal stress with depth and minor horizontal stress (McQueen 2004). The
relationships σH = 1.5 + 2σv and σH = 2.5 + 2.3σv are shown in the top graph. The relationship σh = 0.7σH is shown in the
bottom graph.

148
10 | Tunnels in Hawkesbury Sandstone Case Studies

The nine case studies are: undermined by longwall coal mining. Good quality
geological and rock mass behaviour mapping exists
• Upper Canal Tunnels for them.
„„  ataract (1) – excavated in the 1880s by drill
C
& blast, 3.2 m wide x 2.6 m high, horseshoe 10.2.1 Geology
shaped, maximum depth of 70 m The topography of the Cataract and Devines tunnels’
Devines (2) – excavated in 1880s by drill & blast, area shows steep gradients, up to 45°, between the
„„
2.85 m wide x 2.5m high, horseshoe shaped, tunnels and the nearby Nepean River. The tunnels are
maximum depth of 17 m located 200 m from the Nepean River at the closest
point and lie approximately 80 m above the base of
the valley.
• Malabar Outfall Tunnel (3) The tunnels are aligned north-south and are within
„„ Excavated between 1986 - 1990 by roadheader
the upper levels of the Hawkesbury Sandstone. In this
„„ 644 m long 1:4 decline, 5.3 m wide x 5.0 m high, area, the Hawkesbury Sandstone is approximately 190 m
arched roof, up to 175m depth, although the roof thick comprising sub-horizontal beds 0.5 to 4 m thick.
failure occurred at 100 m depth
Joints are typically 2 to 10 m spaced and are near
vertical. The orientations of the joint sets are shown in
• Sydney LPG Storage Cavern (4) the stereoplot in Figure 10.3 and are consistent with
„„ Excavated between 1996 – 2000 by drill & blast stress relief associated with the Nepean River valley.
„„ Four parallel galleries each 230 m long, 11 m The major joint set is oriented parallel to the tunnel
wide x 14 m high, arched roofs, 124 m depth alignment and the minor joint set perpendicular to the
tunnel alignment (Figure 10.3).
• Northside Storage Tunnel
„„  ree TBM drives excavated between 1998 –
Th 10.2.2 Cataract
2000 The unlined Cataract Tunnel carries water for
ww 7 km of 3.8 m diameter, depth 80 m (5)
approximately 3 km from Broughton’s Pass Weir
ww 9 km of 6.5 m diameter, depth 60 m (6)
pondage to the start of the open section of the Upper
Canal at Brooks Point, approximately 4 km west of
ww 4 km of 6.0 m diameter, depth 80 m (7)
Appin (Figure 10.2). The tunnel has a maximum
capacity of 700 Ml/day however it normally
• Cross City Tunnel (8) operates at a flow rate of approximately 250 Ml/day
„„ Excavated between 2003 – 2005 by roadheader (McQueen 2000).
„„ 2 km long ventilation tunnel, 5.1 m wide x 5.5 m The tunnel was excavated by drill & blast in a
high, very slightly curved roof, 58 m depth traditional horseshoe shape with an average height of
2.6 m and width of 3.2 m. The tunnel was excavated
• Lane Cove Tunnel (9) through Hawkesbury Sandstone with a maximum
„„ Excavated between 2004 – 2006 by roadheader overburden depth of approximately 70 m and was
„„ 2.1 km long road tunnel of 9 m and 12.5 m totally unsupported.
width x 6 m high, very slightly curved roof, 20 Regular inspections by the government agency
- 40 m depth. Sydney Water and its predecessors and advisors had
been carried out in 1961, 1971, 1978, 1985, 1987,
1990 and 1996. More frequent inspections have been
10.2 UPPER CANAL TUNNELS undertaken since 1996 as longwall coal mining is
The Upper Canal is part of the Upper Nepean Scheme carried out underneath the tunnel. The mapping and
which was completed in 1888 to bring water 64 km observations from the 1996 inspection are reported
from the Cataract, Avon, Cordeaux and Nepean in McQueen (2000). Figure 10.4 reproduces the
Dams to the Prospect Reservoir, Sydney. It is critical geological long section that was prepared from the
to Sydney providing approximately 20%, but at 1996 inspection.
times up to 40%, of its water supply (SCA 2012).
The Upper Canal comprises approximately 44 km of Though the intended tunnel profile was an arched
open channels, 784 m of pipe aqueducts and 19 km of roof as seen in the photographs in Figure 10.9, rock
tunnels (Figure 10.2). There are four unlined tunnels fall during and post construction has resulted in a
along the Upper Canal; the Nepean (7190 m long), the roughly square profile with curved shoulders over a
Cataract (2970 m long) and the two Devines tunnels large part of the tunnel (see Figure 10.5 to Figure 10.7
(183 and 817 m long). for examples). Much of this rock fall is inferred to be
the result of stress-induced failures.
The Cataract and Devines tunnels have been
intensely scrutinised since 1996 as they are being The major rock falls in the tunnel have been
experienced in two areas: at around Ch 1350 and 1430 m

149
ESTIMATING ROCK MASS PROPERTIES

FIGURE 10.2 Overall map of the Upper Canal showing the relative locations of the Cataract and Devines tunnels (SCA 2012).

150
10 | Tunnels in Hawkesbury Sandstone Case Studies

FIGURE 10.3 Lower hemisphere polar stereonet of mapped joints in the Devines tunnels (Swarbrick 2013)

(Figure 10.4). In these two areas, the overburden is 50 with values of σ1 = 5.2 MPa oriented at 050° at 48 m
to 70 m deep. Less developed stress fractures were also depth in interbedded siltstone and 8.1 MPa oriented
observed in sections of the tunnel beneath shallower at 120° at 60 m depth in stiffer, massive sandstone
overburden (McQueen 2000). In all cases the stress- (McQueen 2000). The magnitude of the horizontal
induced failures and fractures have occurred within stress is higher than the typical stress values in Sydney
massive sandstone beds which are immediately below (Bertuzzi 2014; Bertuzzi 2014; Pells 2004) and
low strength bedding planes or low stiffness siltstone probably reflects the stiffer sandstone particularly at 60
layers in the roof (e.g. Figure 10.5 to Figure 10.7). m depth. The stress measurements were not necessarily
Stress measurements by over-coring in 1998 carried out in the failure areas.
indicated that the major principal stress σ1 is horizontal

151
ESTIMATING ROCK MASS PROPERTIES

FIGURE 10.4 Geological long section of Cataract Tunnel (McQueen 2000)

152
10 | Tunnels in Hawkesbury Sandstone Case Studies

FIGURE 10.5 Overbreak in the roof of the Cataract tunnel. Sandstone blocks have fallen back to micaceous laminations
due to stress fracturing (SWC-Transwater 1996)

FIGURE 10.6 Cataract tunnel. Inverted V-shaped roof failure in massive sandstone below siltstone. Rock bolts were
installed when the tunnel was undermined by longwall coal extraction (McQueen 2000)

153
ESTIMATING ROCK MASS PROPERTIES

FIGURE 10.7 Cataract tunnel. Roof failure and fallout due to lateral compressive stress in massive sandstone below
interbedded siltstone (McQueen 2000)

154
10 | Tunnels in Hawkesbury Sandstone Case Studies

10.2.3 Devines An inspection in April 1996 (AWT 1998) found:


Two tunnels are collectively known as Devines Tunnel: • e exposed surfaces in the tunnels were in good

Th
Devines Tunnel No. 1 (183 m long) and Devines Tunnel condition and only slightly weathered in most places
No. 2 (817 m long). The two tunnels are separated by • A continuous shale lens 0.5 to 1 m thick lies
a short length of open channel (Figure 10.8). The approximately 3 to 4 m above the tunnels’ roofs.
tunnel sections have a horizontal floor, vertical sides The location of this lens appeared to be higher closer
and an arched crown. The tunnels are typically 2.85 to the southern portal
m wide and 2.5 m high. The longitudinal slope of the • There are zones of weathered silty seams and blast
tunnel invert is approximately 0.4% and both tunnels induced fractures in the roofs of the tunnels
are unsupported and unlined. Both tunnels are very
shallow, with the depth of cover over the tunnels
• Fine grained beds (siltstone, shale and laminite)
within the tunnels have deteriorated
ranging from approximately 7.5 to 12.5 m in Devines
Tunnel No. 1, and approximately 6 to 17 m in Devines • Drummy rock is present in the roof and walls
Tunnel No. 2 (Figure 10.9). • The presence of flaky rock fragments in the roofs
(often in micaceous sandstone beds)
• There are minor sections of the roofs and shoulders
of the tunnels where blocks have the potential to
dislodge at some unknown time in the future.

FIGURE 10.8 Inferred geological long section (top) and topographical cross section (bottom) of Devines Tunnel
(Swarbrick 2013)

155
ESTIMATING ROCK MASS PROPERTIES

FIGURE 10.9 Devines Tunnel No. 1 Southern Portal (left) and Devines Tunnel No. 2 Southern Portal (right) (Swarbrick 2013)

10.2.4 Numerical modelling


The geometry of the Cataract and Devines tunnels Cataract tunnel is shown in Figure 10.10. Three
and the mapped geotechnical conditions where the sections were analysed for the Devines tunnels, Ch
observed stress induced failures and fractures occurred 140, 150 and 595 m. The model at Ch 150 m is shown
were modelled in the two-dimensional finite element in Figure 10.11.
program Phase2 (11). The objective of the modelling The finite element modelling was carried out with
was to assess the likely magnitudes of the major and σh = σEW = 1.5 + 2σv acting in the cross-section of the
minor principal stresses (σ1 and σ3) in the plane of tunnels which is consistent with the over-coring results
the tunnels’ cross-sections, i.e. perpendicular to the obtained at 48 m depth within the Cataract tunnel.
tunnels’ alignments, which were present at the time The effect of including bedding planes in the finite
the failures occurred in the tunnels. element models was assessed by comparing the results
Eight sections in the Cataract tunnel where failure of models of the Cataract tunnel at Ch 460 m with
was recorded were analysed: chainages (Ch) 460, 500, and without bedding planes.
680, 850, 1200, 1408, 2008 and 2490 m. An example Figure 10.12 for Cataract tunnel and Figure 10.13
of the geometry of the finite element mesh used for for Devines tunnels are examples.

156
10 | Tunnels in Hawkesbury Sandstone Case Studies

FIGURE 10.10 Geometry of mesh used in Phase2 finite element analysis for the Cataract tunnel (this example is for Ch 460 m)

157
ESTIMATING ROCK MASS PROPERTIES

FIGURE 10.11 Geometry of mesh used in Phase2 finite element analysis for the Devines tunnels (No 1 tunnel Ch 150 m)

158
10 | Tunnels in Hawkesbury Sandstone Case Studies

FIGURE 10.12 Major (left) and minor (right) principal stresses predicted around the Cataract tunnel (this example is for
Ch 460 m)

FIGURE 10.13 Major (left) and minor (right) principal stresses predicted around the Devines tunnel (No 1 tunnel Ch 150 m)

The inclusion of bedding planes into the finite


element model does affect the predicted stresses as
shown in Figure 10.14. Hence, in cases where the
bedding planes have been mapped, these will be
included in the numerical models.

159
ESTIMATING ROCK MASS PROPERTIES

10

6
σ1 (MPa)

2 Rock mass: sc=30MPa, m=4.11, s=0.0357, a=0.501

Failed

Stable - Close
1
Failed - No bedding

Stable - close, no bedding

0
-1 0 1 2

σ3 (MPa)

FIGURE 10.14 Principal stress plot for the Cataract tunnel Ch 460 m for two models; one without bedding included (green
triangles) and one with bedding included (blue diamonds). Points within the failed area are represented by solid symbols and
those in adjacent stable areas by open symbols.

160
10 | Tunnels in Hawkesbury Sandstone Case Studies

10.3 MALABAR McQueen (2000) describes stress induced failure


occurring in the outfall tunnel within:
10.3.1 Description • The Newport Formation at the bottom station
The Malabar ocean sewerage outfall tunnel is situated • The TBM drive (but more commonly experienced
on the coast to the southeast of Sydney’s CBD (Figure as stress relief movement along bedding planes)
10.15). The outfall tunnel comprises: • The Hawkesbury Sandstone in the road header
• A 1:4 decline, 5.3 m wide by 5.0 m high and 644 decline.
m long, excavated by roadheader and drill & blast The stress failure in the decline within the
within the Hawkesbury Sandstone and underlying Hawkesbury Sandstone occurred more than a year
sandstones of the Newport Formation after excavation. The sandstone fractured up to a
• A bottom station, 6.7m wide by 5.7 m high and siltstone bed approximately 0.3m above the roof at an
150 m long, excavated in the Newport Formation overburden depth of 100 m (see Figure 10.16).
at -155 mRL (below about 175 m of overburden) by The in situ stress was measured in the Malabar
drill & blast tunnel within the Newport Formation sandstone
• A 4 m diameter, 3.9 km long TBM excavated which underlies the Hawkesbury Sandstone. It
tunnel in the Newport Formation and underlying confirms the typical in situ stresses found in Sydney as
Bald Hill claystone and Bulgo Sandstone. shown in Figure 10.1.
The outfall tunnel was constructed between 1986
and 1990.

FIGURE 10.15 Location of case studies around Sydney’s CBD – Northside Storage Tunnel, Lane Cove Tunnel in the north, Cross City
Tunnel in the centre, and Malabar outfall tunnel and LPG Storage Cavern in the south. Base map sourced on 3rd January 2014.

161
ESTIMATING ROCK MASS PROPERTIES

10.3.2 Numerical modelling The objective of the modelling is to assess the likely
The geometry of the outfall tunnel decline and magnitudes of σ1 and σ3, which were present at the
the mapped geotechnical conditions where the time of the rock fall.
rock fall occurred were modelled in the two- The geometry of the finite element mesh used is
dimensional finite element program Phase2 . shown in Figure 10.17.

FIGURE 10.16 Malabar outfall tunnel, decline roof failure in Hawkesbury Sandstone up to a laminated siltstone bed. The
rock bolts were installed after the fallout which occurred a year after excavation (McQueen 2000)

162
10 | Tunnels in Hawkesbury Sandstone Case Studies

FIGURE 10.17 Geometry of mesh used in Phase2 finite element analysis for the decline section of the Malabar ocean outfall
tunnel

The in situ stress was modelled as that typical of


Sydney, such that the horizontal stress acting in the
cross section of the modelled east-west oriented tunnel
is σH = σNS = 1.5 + 2σv . At a depth of 100 m, this gives
σH = 6.5 MPa, which is consistent with the measured
σH as shown in Figure 10.1.
The results of this modelling are presented in the
contour plots in Figure 10.18.

163
ESTIMATING ROCK MASS PROPERTIES

FIGURE 10.18 Contours of major (top) and minor (bottom) principal stresses from finite element analysis of the decline
section of the Malabar ocean outfall tunnel. The extent of the mapped spalling is shown.

164
10 | Tunnels in Hawkesbury Sandstone Case Studies

10.4 SYDNEY LPG STORAGE Two large roof falls occurred during the excavation
of the galleries, which was carried out by drill and
CAVERN blast methods. The failures occurred in Galleries A
and C. de Ambrosis and Kotze (2004) provide a good
10.4.1 Description description of the failures.
An underground facility to store 65,000 tonnes of
Figure 10.20 shows the mapped details on a cross-
liquefied petroleum gas (LPG) was constructed at
section of the first roof failure which occurred in the
Molineux Point, Port Botany, Sydney between 1996
Gallery A heading which was driven as a 6 m wide
and 2000 (see Figure 10.15). The storage facility
central pilot heading followed by the removal of 2.5 m
comprises four parallel galleries, each approximately
wide side strips.
230 m long and rectangular 14 m high and 11 m
wide, with arched roofs. The galleries are at a depth The failure occurred during or immediately after
of 124 m below ground surface and are totally within the perimeter blast of the western strip at Ch 096 m
the Hawkesbury Sandstone. They are approximately when the pilot heading was at Ch 104 m. The collapse
aligned with their long axis running north-south. was over a 13 m length between Ch 089 to 102 m,
across 7 m wide roof span and extended 1 to 1.5m into
The LPG is contained by the pressure of the
the roof (Figure 10.20). It is understood that at the
surrounding groundwater. A water curtain tunnel and
time of failure, roof support was provided only by end-
a network of radiating water filled injection drill holes
anchored, 4 m long, hollow rock bolts on a 2 x 2 m
is located 15 m above the storage galleries to ensure
grid pattern (de Ambrosis and Kotze 2004).
saturation of the rock (see Figure 10.19).

FIGURE 10.19 Isometric view of the LPG storage cavern; from de Ambrosis and Kotze (2004). Gallery A is the closest to the
access shaft (leftmost in this view), then sequentially labelled Gallery B, C and D.

165
ESTIMATING ROCK MASS PROPERTIES

FIGURE 10.20 Failure in Gallery A; from de Ambrosis and Kotze (2004)

The second failure occurred in Gallery C which was 10.4.2 Numerical modelling
supported by 5 m long, fully cement grouted, hollow The geometry of the cavern and the mapped
rock bolts on a 1.5 x 2 m grid pattern. The failure geotechnical conditions where these failures occurred
occurred approximately 7.5 hours after blasting over a were modelled in the 2D finite element program
7 m length between Ch 031 to 038m and extended 0.8 Phase2 . The objective of the modelling is again to assess
to 1.2 m into the Hawkesbury Sandstone roof. the likely magnitudes of σ1 and σ3, which were present
at the time of failure.
Figure 10.20 shows the characteristics of the
Hawkesbury Sandstone at the failures which The geometry of the finite element mesh used is
comprised: shown in Figure 10.21.
Three cases of in situ stress were modelled.
• Fresh, strong, medium to coarse grained, sheet
facies sandstone First, that which is typical of Sydney, such that
• The concave-shaped cross beds within the sandstone the horizontal stress acting in the cross section of the
dipped at 12 to 22° to the horizontal modelled east-west oriented tunnel is σh = σEW = 0.7 x
(1.5 + 2σv). At 124 m depth, that is σh = 5.4 MPa.
• Beds from 0.3 to 2.5 m thick with sub-horizontal,
planar to undulating bedding planes some thinly The second case was modelled as the inferred failures
coated with carbonaceous silt and clayey sand. were initiated by ‘compression induced movement’ (de
de Ambrosis and Kotze (2004) inferred that the roof Ambrosis and Kotze 2004), which may imply high
strata were initially fractured by compression induced horizontal stress. To test this case, models were run
movement along the bedding planes towards the centre with the in-plane horizontal stress = 1.5 + 2σv.
of the heading. Collapse of the roof then occurred as The third case analysed was hydrostatic stress, in
a result of bending induced tension in the sandstone. keeping with the unpublished reports in McQueen (2004).
The weight of the fractured rock being too great for the In this case at 124 m depth, σh = 3.1 MPa. It is noted that
installed rock bolts. the data point used to represent the major principal stress
In situ stress measurements from over-coring (Figure 10.1) lies between the modelled cases.
have not been published but were reported to be The results of this modelling are presented as
less than the typical base case shown in Figure 10.1 contour plots of major and minor principal stresses in
(McQueen 2004). Figure 10.22.

166
10 | Tunnels in Hawkesbury Sandstone Case Studies

FIGURE 10.21 Geometry of mesh used in Phase2 finite element analysis for the LPG storage caverns

167
ESTIMATING ROCK MASS PROPERTIES

FIGURE 10.22 Contours of major (top) and minor (bottom) principal stresses from finite element analysis of the first roof
collapse with the base case in situ stress. The extent of the collpase as observed is shown.

168
10 | Tunnels in Hawkesbury Sandstone Case Studies

10.5 NORTHSIDE STORAGE „„  owever, the impact of spalling in the 6.0 m


H
diameter Tunks Park to Scott’s Creek drive was
TUNNEL more serious as its TBM was a shielded machine
compared to the open TBM used in the Tunks
10.5.1 Description Park to North Head tunnel. The shielded machine
The Northside Storage Tunnel (NST) was designed meant the first opportunity to install rock bolts
and constructed between 1998 and 2000 to capture was about 15 m from the face (i.e. > 2 x tunnel
and detain the major storm overflows from an existing diameter), compared with approximately 6 m in
sewer to reduce the raw sewage entering and polluting the open machine (i.e. ≈ 1 x tunnel diameter).
Sydney Harbour. It comprises 20 km of TBM driven The overburden depth in Tunks Park to Scott’s
tunnels, 3 km of declines and caverns (Asche and Creek drive was approximately 80 m.
Quigley 1999). „„ 
Some spalling also occurred at the start of the
The 20 km of TBM drives comprised the following 3.8 m diameter Tunks Park to Lane Cove TBM,
which were all excavated within the Hawkesbury also at a depth of approximately 80 m.
Sandstone (Figure 10.15 and Figure 10.23): The typical in situ stress shown in Figure 10.1 is also
• 7 km of 3.8 m diameter (Tunks Park to Lane Cove) the line of best fit of the numerous hydrofracture data
• 9 km of 6.5 m diameter (Tunks Park to North Head) available from this project, ignoring those specific data
affected by topography (McQueen 2004).
• 4 km of 6.0 m diameter (Tunks Park to Scott’s Creek)
Several failures occurred in the TBM drives during
construction (Gee et al. 2002).
10.5.2 Numerical modelling
The geometry of the tunnels and the inferred
• A roof fall occurred several months after the TBM geotechnical conditions where the spalling failures
tunnel was excavated in the Manly area (Tunks Park occurred in the main east-west drives from Tunks
to North Head). The tunnel passed underneath a Park were modelled in Phase2 . The objective of the
palaeochannel and it was considered by Gee et al. modelling was again to assess the likely magnitudes
(2002) that the fall was caused by groundwater of σ1 and σ3 which had occurred in the spalling zone.
penetrating the bedding plane of a shale unit just
above the tunnel crown. The geometry of the finite element mesh used is
shown in Figure 10.24.
• Spalling caused significant problems in the main
east-west drives from Tunks Park. Breakage in The in situ stress was modelled as that typical of
the invert caused problems with the supply trains’ Sydney, i.e. σH = σNS = 1.5 + 2σv and σh = 0.7σH. This
tracks and in the crown which necessitated more is consistent with the project data (Figure 10.1).
bolting and mesh. The results of this modelling are presented as a series
„„ 
The first TBM drive to encounter spalling was of contour plots in Figure 10.25 to Figure 10.27.
the 6.5 m diameter Tunks Park to North Head
tunnel when the tunnel was at approximately 60
m depth. Spalling occurred within 60 minutes of
excavation (Wallis 2000).

FIGURE 10.23 Alignment of the NST; from Gee et al. (2002)

169
ESTIMATING ROCK MASS PROPERTIES

FIGURE 10.24 Geometry of mesh used in Phase2 finite element analysis for the 6.5 m diameter Tunk’s Park to North Head
tunnel

170
10 | Tunnels in Hawkesbury Sandstone Case Studies

FIGURE 10.25 Contours of major (top) and minor (bottom) principal stresses from finite element analysis of the 6.5 m
diameter Tunks Park to North Head drive under 60 m overburden. The extent of the inferred spalling may be seen.

171
ESTIMATING ROCK MASS PROPERTIES

FIGURE 10.26 Contours of major (top) and minor (bottom) principal stresses from finite element analysis of the 6.5 m
diameter Tunks Park to North Head drive under 95 m overburden. The extent of the inferred spalling may be seen.

172
10 | Tunnels in Hawkesbury Sandstone Case Studies

FIGURE 10.27 Contours of major (top) and minor (bottom) principal stresses from finite element analysis of the Lane Cove,
3.8m diameter Tunks Park to Lane Cove drive under 80 m overburden. The extent of the inferred spalling may be seen.

173
ESTIMATING ROCK MASS PROPERTIES

10.6 CROSS CITY TUNNEL the tunnellers, Ronald Shores. The rock fall occurred
close to the tunnel’s advancing face at a point where
10.6.1 Description the tunnel was just rising out of its low point. Along
The Cross City Tunnel (CCT) provides a 2.1 km long this section, the ventilation tunnel is 5.1 m wide x 5.5
east-west connection through Sydney’s CBD between m high at a depth of 58 m, with a very slightly curved
Darling Harbour in the west and Rushcutters Bay in crown (PSM 2005) (13).
the east (Figure 10.28 and Figure 10.29). The CCT Mapping at the time identified the following
comprises twin 2-lane road tunnels, associated ramp features (Figure 10.30):
tunnels and a ventilation tunnel. The road tunnels • Ellipsoidal shaped fallout slightly to the north (left)
follow separate alignments over the western half. The of tunnel centre
CCT was constructed between January 2003 and • Failure was in a relatively thin layer of light grey
August 2005. massive sandstone
A rock fall occurred during the construction of the • The perimeter of the failure area appeared to have
ventilation tunnel on the 29 July 2004 killing one of the characteristics of tensile failure.

FIGURE 10.28 Alignment of the CCT; from Battaglia et al. (2006)

FIGURE 10.29 Schematic long section showing the relative position of the main road tunnels and the deeper ventilation
tunnel; from Battaglia et al. (2006)

(13)
The author and his colleagues from PSM were brought in to assist the investigation

174
10 | Tunnels in Hawkesbury Sandstone Case Studies

FIGURE 10.30 Mapping of geology at the failure by GHD taken from PSM (2005)

10.6.2 Numerical modelling


The geometry of the ventilation tunnel and the mapped The geometry of the finite element mesh used is
geotechnical conditions where the rock fall occurred shown in Figure 10.31.
were modelled in the two-dimensional finite element The in situ stress was modelled as that typical of
program Phase2 . The objective of the modelling is to Sydney, i.e. σH = σNS = 1.5 + 2σv acting in the plane of
assess the likely magnitudes of the major and minor the tunnel cross-section, and σh = 0.7σH.
principal stresses, σ1 and σ3, which were present at the The results of this modelling are presented in the
time of the rock fall. contour plots in Figure 10.32.

FIGURE 10.31 Geometry of mesh used in Phase2 finite element analysis for the CCT ventilation tunnel

175
ESTIMATING ROCK MASS PROPERTIES

FIGURE 10.32 Contours of major (top) and minor (bottom) principal stresses from finite element analysis of the
ventilation tunnel

176
10 | Tunnels in Hawkesbury Sandstone Case Studies

10.7 LANE COVE TUNNEL either massive or cross bedded at dips of 20 to 30°. Parts
of the associated ramp tunnels, ventilation tunnels also
10.7.1 Description encountered the overlying Mittagong Formation and
The Lane Cove Tunnel (LCT) is a 3.6 km long Ashfield Shale. Bedding was sub-horizontal, typically
underground road link between the Gore Hill Freeway dipping at 0-5° towards the south. Joint sets are sub-
at Artarmon and the M2 motorway at Lane Cove vertical and trend approximately north-south and east-
River (Figure 10.33). The LCT comprises twin two west (Badelow et al. 2005).
and three-lane road tunnels, associated ramp tunnels
and ventilation tunnels (Badelow et al. 2005). The 10.7.2 Numerical modelling
LCT was constructed between January March 2004 The geometry of the LCT tunnels and the mapped
and December 2006. geotechnical conditions where the rock fall occurred
The intersection of a ventilation tunnel and ramp were modelled in the two-dimensional finite element
tunnel, which was designed by others collapsed at program Phase2 . Again, the objective of the modelling
1:40am on Wednesday 2nd November 2005. This was to assess the likely magnitudes of σ1 and σ3, which
event has been reported elsewhere, e.g. Wallis (2007). were present at the time of the rock fall.
It is not a stress induced failure and is not part of The specific location where the spalling occurred
this work. was beneath a minor valley associated with the Stringy
The 54 m2 two-lane tunnel has a typical span of 9 Bark Creek. The geometry of the finite element mesh
m, expanding to 13.5 m in the break-down bays. The used is shown in Figure 10.35.
75 m2 three-lane tunnel has a 12.5 m span and 17.5 m The erosional development of the creek valley was
in its breakdown bays. simulated in the model with a single initialisation step
Drummy and spalling shotcrete and minor spalling as indicated in Figure 10.35. The in situ stress was
of Hawkesbury Sandstone in the roofs of the two- and modelled as that typical of Sydney, i.e. σH = σNS = 1.5
three-lane tunnels occurred when the underlying 9 + 2σv acting in the plane of the modelled tunnel cross-
m span ventilation tunnel and a deep services trench section and σh = 0.7σH.
within the road tunnels were excavated (Clark 2005). The resultant major horizontal stress in the crown
The main road tunnels of the LCT were excavated of the LCT tunnels is in the order of 3MPa (Figure
within the Hawkesbury Sandstone which is composed 10.34 which is in keeping with the high horizontal
predominantly of fine to medium grained quartzose stresses associated with topographic relief as suggested
sandstone deposited in 1 to 3 m thick beds which are in Figure 101.

FIGURE 10.33 Alignment of the LCT; from www.whereis.com accessed 16th October 2013

177
ESTIMATING ROCK MASS PROPERTIES

FIGURE 10.34 Predicted major principal stress acting in the horizontal direction in the plane of the modelled cross section
(σH) after the valley is modelled and shows σH of approximately 3.0 MPa in the crown of the main tunnel.

FIGURE 10.35 Geometry of mesh used in Phase2 finite element analysis for the LCT tunnels. In the model the Stringy Bark
Creek valley is first created.

178
10 | Tunnels in Hawkesbury Sandstone Case Studies

The results of this modelling are presented in the


contour plots in Figure 10.36.

FIGURE 10.36 Contours of major (top) and minor (bottom) principal stresses from finite element analysis of the LCT three-
lane tunnel (Ch 1510 m). The extent of the mapped spalling is shown. 

179
ESTIMATING ROCK MASS PROPERTIES

The second is the Hoek-Brown strength criterion for


10.8 RESULTS the Hawkesbury Sandstone Class I /II rock mass. This
The results of the numerical modelling of the nine case has been derived using the GSI approach with GSI =
studies described in the preceding sections are presented 70 (range 65 – 80) (Figure 9.15) which yields mb = 4.11,
in the principal stress plots shown in Figure 10.37 to s = 0.0357 and a = 0.501.
Figure 10.42. In these figures, solid symbols are used
to identify the inferred stresses in “failed zones” and The third strength criterion shown in the figures
open symbols for “stable zones”. The envelopes of four has been suggested to define the onset of spalling
strength criteria are also shown in the plots. as discussed in Chapter 5.2 (Diederichs et al. 2007;
Diederichs et al. 2004). This can be written for
The first is the Hoek-Brown strength criterion for Hawkesbury Sandstone Class I /I as Equation 10.1.
intact Hawkesbury Sandstone Class I /II rock with
parameters mi = 12 ≈ (σc / |σt |) and s = 1. A typical
value of σc is 30 MPa, from Figure 2.1.
However, as discrete blocks of sandstone were not
included in the finite element analyses of the tunnels,
there is an argument that a compressive strength
appropriate for a volume of rock larger than a UCS test EQUATION 10.1
should be used. A nominal volume of a 1 m3 cube of
Hawkesbury Sandstone is adopted. This is in keeping The fourth envelope shown depicts the ‘systematic
with studies for coal pillars as will be further discussed cracking’ or ‘damage initiation’ criteria of Diederichs
in the next chapter. According to Masoumi (2013), the (2007) (Equation 5.2)
reduction in unconfined compressive strength beyond
a diameter of 50 to 65 mm is:


EQUATION 10.2

For Hawkesbury Sandstone Class I /II Equation 10.2 is:


This relationship was based on samples of up to
200 mm. Assuming this relationship continues for
larger sample sizes, then for a 1 m3 cube, σc =30 *
(65/1000)0.11, that is σc = 22.2 MPa. and

180
10 | Tunnels in Hawkesbury Sandstone Case Studies

18

16

14

12

10
σ1 (MPa)

6
1m3: sc=22.2MPa, mi=12

Rock mass: sc=22.2MPa,


m=4.11, s=0.0357, a=0.501
4 Initial spalling: 6.66MPa + 2s3

Damage initiation: sc=22.2MPa,


m=0.0972, s=0.0081, a=0.25
Cataract - stable

2 Devines - stable

CCT - stable

Elgas - stable

0 LCT - stable

-1 0 1 2 3 4
Malabar - stable

NST - stable
σ3 (MPa)

FIGURE 10.37 Principal stress plot showing only the modelling predicted stresses from the ‘stable’ zones compared with
proposed strength envelopes

181
ESTIMATING ROCK MASS PROPERTIES

18

16

14

12

10
σ1 (MPa)

6
1m3: sc=22.2MPa, mi=12

Rock mass: sc=22.2MPa,


m=4.11, s=0.0357, a=0.501
4 Initial spalling: 6.66MPa + 2s3

Damage initiation: sc=22.2MPa,


m=0.0972, s=0.0081, a=0.25
Cataract - stable (close)

Devines - stable (close)


2
CCT - stable (close)

Elgas - stable (close)

LCT - stable (close)


0
-1 0 1 2 3 4
Malabar - stable (close)

NST - stable (close)


σ3 (MPa)

FIGURE 10.38 This is a subset of Figure 10.37 and shows the predicted stresses from the ‘stable’ zones immediately
adjacent to the tunnel boundary compared with proposed strength envelopes

182
10 | Tunnels in Hawkesbury Sandstone Case Studies

18

16

14

12

10
σ1 (MPa)

1m3: sc=22.2MPa, mi=12


8
Rock mass: sc=22.2MPa,
m=4.11, s=0.0357, a=0.501
Initial spalling: 6.66MPa + 2s3

6 Damage initiation: sc=22.2MPa,


m=0.0972, s=0.0081, a=0.25
Cataract - failed

Devines - failed

4 CCT - failed

Elgas - failed

LCT - failed
2
Malabar - failed

NST - failed

0
-1 0 1 2 3 4

σ3 (MPa)

FIGURE 10.39 Principal stress plot showing only the modelling predicted stresses from the ‘failed’ zones compared with
proposed strength envelopes.

183
ESTIMATING ROCK MASS PROPERTIES

18

16

14

12

10
σ1 - σ3 (MPa)

6
1m3: sc=22.2MPa, mi=12

Rock mass: sc=22.2MPa, m=4.11,


s=0.0357, a=0.501
Initial spalling: 6.66MPa + 2s3
4 Damage initiation: sc=22.2MPa,
m=0.0972, s=0.0081, a=0.25
Cataract - stable

Devines - stable

2 CCT - stable

Elgas - stable

LCT - stable

Malabar - stable
0
-1 0 1 2 3 4
NST - stable

σ3 (MPa)

FIGURE 10.40 Deviator stress (σ1-σ3) versus confining stress (σ3) plot showing only the modelling predicted stresses from the
‘stable’ zones compared with proposed strength envelopes.

184
10 | Tunnels in Hawkesbury Sandstone Case Studies

18

16

14

12

10
σ1 - σ3 (MPa)

8 1m3: sc=22.2MPa, mi=12

Rock mass: sc=22.2MPa,


m=4.11, s=0.0357, a=0.501
Initial spalling: 6.66MPa + 2s3
6
Damage initiation: sc=22.2MPa,
m=0.0972, s=0.0081, a=0.25
Cataract - stable (close)

Devines - stable (close)


4
CCT - stable (close)

Elgas - stable (close)

2 LCT - stable (close)

Malabar - stable (close)

NST - stable (close)

0
-1 0 1 2 3 4

σ3 (MPa)

FIGURE 10.41 This is a subset of Figure 10.40 and shows the predicted deviator stresses from the ‘stable’ zones immediately
adjacent to the tunnel boundary compared with proposed strength envelopes

185
ESTIMATING ROCK MASS PROPERTIES

18

16

14

12

10
σ1 - σ3 (MPa)

8 1m3: sc=22.2MPa, mi=12

Rock mass: sc=22.2MPa,


m=4.11, s=0.0357, a=0.501
Initial spalling: 6.66MPa + 2s3
6
Damage initiation: sc=22.2MPa,
m=0.0972, s=0.0081, a=0.25
Cataract - failed

Devines - failed
4
CCT - failed

Elgas - failed

LCT - failed
2
Malabar - failed

NST - failed

0
-1 0 1 2 3 4

σ3 (MPa)

FIGURE 10.42 Deviator stress (σ1-σ3) versus confining stress (σ3) plot showing only the modelling predicted stresses from the
‘failed’ zones compared with proposed strength envelopes.

186
10 | Tunnels in Hawkesbury Sandstone Case Studies

10.9 DISCUSSION 10.10 CONCLUSIONS


In many of the cases, bedding planes were mapped as Nine case studies where stress induced failure occurred
either intersecting or close to the tunnel excavation. in tunnels excavated within Hawkesbury Sandstone
Not surprisingly, one of the results of the analyses is were presented in this chapter. The case studies are from
that the proximity of bedding planes influences the a variety of sources and cover a range of conditions:
induced stresses surrounding the tunnel. • Construction dates from the 1880s to 2006
The inferred confining stress, σ3, of the nine case • Spans from 3.2 to 12.5 m
studies is rather modest ranging only up to 3 MPa. • Depths from 17 to 124 m
Further, at the locations where failures were mapped, • Drill & blast, roadheader and TBM excavation.
the confinement is usually less than 1 MPa. Yet even
Using two-dimensional finite element modelling,
though the failures are essentially within low or zero
the likely principal stresses that were present at the time
confinement, the Hoek-Brown rock mass strength
of the failures were assessed and then compared with
criterion does differentiate between the ‘failed’ and
the Hoek-Brown and DISL criteria. The parameters for
‘stable’ zones relatively well (Figure 10.37 to Figure
these criteria were assessed from the typical range of
10.39). The same data is plotted as deviator stress
values for Hawkesbury Sandstone.
(σ1- σ3) versus confining stress (σ3) in Figure 10.40 to
Figure 10 .42. Not every ‘failed’ stress point plotted above the
criterion and not every ‘stable’ stress point plotted
Figure 10.43 plots the data as the difference between
below the criterion. However, the work presented
the criterion predicted strength and the major principal
in this chapter showed the remarkable ability of
stress from the models (σ1- σ1^) versus confining stress
the criteria with seemingly generic parameters, to
(σ3). This figure may make it easier to see that the
differentiate between most ‘failed’ and ‘stable’ zones.
Hoek-Brown criterion does identify the failed and
stable zones well, but perhaps the criterion suggests The Hoek-Brown criterion parameters for typical
zones should have failed in places where failure was Hawkesbury Sandstone Class I /II are:
not observed. The Hoek-Brown criterion does identify
zones other than those mapped as failing but these tend • σc = 22 MPa (1 m3 cube), mb= 4.11, s = 0.0357 and
to be in the floors of the tunnels, i.e. zones which are a = 0.501
not as readily picked up in mapping as having failed.
While the damage initiation and initial spalling The damage initiation criterion parameters for
criteria do capture the ‘stable’ zones well, many ‘failed’ typical Hawkesbury Sandstone Class I /II are:
zones plot beneath the rock mass strength inferred
from these criteria.
• σc = 22 MPa (1 m3 cube), mb= 0.0972, s = 0.0081
The damage initiation and Hoek-Brown criteria and a = 0.25
are combined into the Damage Initiation Spalling
Limit (DISL) as recommended by Diederichs (2007)
in Figure 10.44 as principal stress and deviator stress
plots. The DISL differentiates between the ‘failed’
from the ‘stable’ zones very well.

187
ESTIMATING ROCK MASS PROPERTIES

FIGURE 10.43 The difference between the criterion predicted strength and the major principal stress from the models
versus confining stress plots. The top series shows the difference from the Hoek-Brown criterion; the bottom shows the
difference from the Damage Inititaion criterion. The ‘failed’ zones are on the left; ‘stable’ zones in the middle and ‘stable –
close’ zones on the right.

188
10 | Tunnels in Hawkesbury Sandstone Case Studies

FIGURE 10.44 Principal stress (top) and deviator stress (bottom) versus confining stress plots showing the ‘failed’ zones
(left) and ‘stable’ zones (right).

189
ESTIMATING ROCK MASS PROPERTIES

190
11 | Coal Pillar Case Studies

11 COAL PILLAR
CASE STUDIES

11.1 OUTLINE Northern hemisphere coals were formed in


conditions conducive to the formation of coals rich in
Underground coal mining constructs hundreds of
vitrinite and low in mineral matter content. In contrast,
thousands of coal pillars annually. The industry is
the Permian swamps of the southern hemisphere
dependent upon the stability of these variously sized
existed in cold conditions related to the ending of a
hexahedrons so knowledge of the coal’s rock mass
massive ice age. Slow growth allowed the intrusion of
strength is imperative. Yet the approach for estimating
fine inorganic material, present today as the high-ash
the rock mass strength of coal has developed along a
component of the coal. So, there is a broad difference
different path to that for other rock masses.
in coals from the northern and southern hemispheres,
The typical method for estimating rock mass the latter being typically of higher ash content.
strength, which we have discussed in preceding
There are also differences between coals from the
chapters, is to basically reduce the intact rock strength
same geological age. South Africa is noted for its high
by an empirical factor that accounts for the geological
ash, generally lower sulphur, and lower-energy thermal
structures in the rock mass (bedding planes, joints,
coals. The dominant product from underground coal
shears, etc.). The empirical Hoek-Brown / GSI process
mining in New South Wales, Australia is a soft-coking,
is one such method. While the most used methods
high-volatile coal (Tinsley 1982). Indian coal has high
for estimating the strength of coal pillars are also
ash content, medium to high volatile matter and low
empirical they are geometrically based. They do not
calorific value (Mishra 2009). At the outset therefore,
explicitly include either the intact strength of the coal,
it would be reasonable to expect each region to exhibit
the in situ stress, or the geological structures of the
different coal properties based on the genesis of
coal rock mass. These geometrically based methods
the deposits.
are seemingly at odds with the common methods of
estimating rock mass strength. The published data relating to coal pillars comes
largely from Australia, South Africa, the USA and
The approach to estimate coal pillar strength is
India.
examined in this chapter. The case studies of collapsed
and ‘un-collapsed’ coal pillars from Australia, South
Africa, USA, India and elsewhere is first presented. 11.2.1 Australia
Numerical analyses are then carried out with a view A series of mine design accidents in the late 1980s
to assess the compatibility between the analyses and resulted in a research program at the University of
the empirical methods. The combined database is then NSW Australia (UNSW Australia or UNSW) aimed
assessed in relation to the brittle (damage initiation) and at developing pillar design guidelines (Galvin et al.
Hoek-Brown rock mass strength criteria. Appropriate 1999). A database from Australian underground coal
strength parameters for modelling coal pillars are mining was compiled, as reproduced in Table 11.2 and
then proposed. summarised in Table 11.1 and in Figure 11.1.
The database comprises 18 failed (also termed
‘collapsed’) and 15 stable (also termed ‘un-collapsed’
11.2 DATA or ‘intact’) coal pillars (Galvin 1995).
Although coal is known in most geological periods, Based on this database and the earlier work of
the bulk of the world’s coal deposits were formed Salamon and Munro (1967) in South Africa, Galvin
during two main periods of geological history: the et al. (1999) developed empirical relationships to
Carboniferous in the northern hemisphere about 300 estimate coal pillar strength, Cp, as a function of pillar
million years ago, and the Permian in the southern geometry, w/h. The linear version of those relationships
hemisphere about 50 million years later (Cook 1994). for the combined Australian and South African

191
ESTIMATING ROCK MASS PROPERTIES

database is presented in Equation 11.1 (see also Though the relationships are said to be applicable
Equation 5.11). for w/h ≤ 8, only one of the collapsed pillars in the
Australian database has a ratio w/h = 8; the others
having w/h < 5 (see Figure 11.1). Galvin (2006) argues
the validity of the w/h = 8 case, though others have
EQUATION 11.1 questioned it; see Colwell (2010); Seedsman (2012).

TABLE 11.1 Summary of the UNSW Coal Pillar Database


Case Number of Depth [m] w/h
cases

Uncontrolled 1.1 – 3.0


Collapsed 18 58 – 336
Controlled 4.5 – 8.0

Un-collapsed 15 22.5 – 510 2 – 11.2

w/h = R
0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0
0

50

100

150

200

250
Depth [m]

300

350

400

450

500
Uncollapsed
550
Collapsed
600

FIGURE 11.1 The database of Australian coal pillars presented by (Galvin et al. 1999)

192
11 | Coal Pillar Case Studies

TABLE 11.2 Details of the UNSW Coal Pillar Database: the one collapsed case (SC3) of w/h > 5 is
highlighted
Case Depth Pillar Dimensions Pillar w/h Bord Age at
[m] [m] Height Width collapse
(minimum width) [m] [m] [years]

FC1 140 15 x 15 15.0 5.0 3.0 7.0 2.0

FC2 80 7.5 x 18 7.5 7.0 1.07 7.0 0.5

FC3 80 6 x 7.5 6.0 3.0 2.0 7.0 0.5

FC4 135 7 x 20 7.0 5.0 1.4 6.0 0.5

FC5 100 10 x 10 10.0 6.0 1.66 5.5 1.0

FC6 120 10 x 20 10.0 4.5 2.22 7.0 18.0

FC7 95 3.5 x 12 3.5 1.75 2.0 6.2 -


Collapsed

FC8 70 9x9 9.0 6.0 1.5 8.0 0.3

FC9 250 15 x 27 15.0 3.0 5.0 15.0 0.5

FC10 115 10.5 x 22 10.5 5.0 2.1 8.0 0.3

FC11 145 11.25 x 22.5 11.25 7.0 1.6 6.25 4.0

FC12 336 13.5 x 28.5 13.5 3.0 4.5 6.3 0.5

FC13 185 21 x 21 21.0 4.9 4.28 6.1 -

FC14 240 22 x 22 22.0 4.9 4.48 5.1 -

FC15 152 25 x 30 20.0 9.2 2.17 5.5 10.0

SC1 75 8x8 8.0 4.5 1.77 6.5 1.0

SC2 58 3.65 x 54 3.65 2.7 1.35 7.3 1.0

SC3 170 20 x 20 20.0 2.5 8.0 5.5 0.3

FS1 80 23.5 x 23.5 23.5 2.1 11.19 7 33

FS2 122 7.6 x 30.5 7.6 1.5 5.07 6 85

FS3 40 10 x 10 10.0 3.2 3.13 6 1.5

FS4 22.5 2x4 2.0 1.0 2 5 170

FS5 58 5 x 30 5.0 1.7 2.94 6 85


Un-collapsed

FS6 70 14 x 14 14.0 2.5 5.6 6.5 43

FS7 510 23.5 x 44.5 23.5 2.9 8.23 7 5

FS8 500 23.5 x 44.5 23.5 2.8 8.56 7 5

FS9 470 23.5 x 44.5 23.5 2.7 8.87 7 5

FS10 450 23.5 x 44.5 23.5 2.6 3.22 7 5

FS11 150 16 x 18 16.0 3.5 4.6 7 30

FS12 75 4 x 11 4.0 1.0 4 5 40

FS13 50 10 x 10 10.0 6.0 1.67 5.5 4

FS14 430 32 x 36 32.0 3.0 10.66 6 -

FS15 75 15 x 15 15.0 3.0 5 15 6

193
ESTIMATING ROCK MASS PROPERTIES

11.2.2 South Africa It is important to appreciate that no pillar failure


The most recent database of South African coal pillar has been recorded in South African collieries with a
experience is that collated by van der Merwe and w/h ratio greater than 4.45 (Salamon et al. 2006). That
Mathey (2013), which is summarised in Table 11.3 is, the maximum w/h for collapsed pillars is 4.45. Also,
and in Figure 11.2 as a plot of depth against the ratio the maximum depth in the South African database
w/h. The individual cases are reproduced in Table 11.4 is 255 m.
for the 86 collapsed cases and Table 11.5 for the 337 The empirical relationship to estimate coal pillar
un-collapsed cases. strength, Cp, given by van der Merwe (2003) for

TABLE 11.3 Summary of the South African coal pillar database


Case Number of Depth [m] w/h
cases

Normal coal 65 19 – 205 0.9 – 3.1


Collapsed
Weak coal 21 58 – 152 1.0 – 4.4

Un-collapsed 337 13 – 255 1.0 – 16.2

w/h
0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0
0

50

100

150

200

250
Depth [m]

300

350

400

450

500
Intact

550 Collapsed
Weak coal Collapsed
600

FIGURE 11.2 The database of South African coal pillars presented by van der Merwe and Mathey (2013)

194
11 | Coal Pillar Case Studies

‘normal coal’ is reproduced in Equation 11.2 (from No relationship for ‘weak coal’ was provided by
Equation 5.12). van der Merwe (2003) though in earlier work, van der
Merwe (1993) did suggest the constant in the Salamon
and Munro (1967) equation (Equation 5.3) should
be 4.5 MPa for ‘weak’ coal, compared to 7.2 MPa
for ‘normal’ coal. Interestingly, Mathey and van der
Merwe (2015) from their analysis of 1800 laboratory
tests found no significant differences in strength
EQUATION 11.2
properties between South African coal seams.

TABLE 11.4 Details of the South African Coal Pillar Database – Collapsed cases
Pillar Bord Mining Tributary
Depth
Case Coalfield width width height w/h Load
[m]
[m] [m] [m] [MPa]

n201 Free State 29 5.4 6.3 2.9 2.2 3.4

n202 Free State 60 7 6 1.82 3.3 5.2

n203 Free State 53 5.6 6.1 1.8 3.4 5.8

n204 Free State 21 6.75 5.25 3.2 1.6 1.7

n205 Free State 19 6 6 3 2.0 1.9

n206 Free State 23 6 6 2.9 2.1 2.3

n207 Free State 40 7 5 2.8 1.8 2.9

n208 Free State 42 4.5 5.5 2 2.8 5.2

n209 Free State 50 7 5 2.3 2.2 3.7

n210 Free State 55 7 5 2.2 2.3 4.0

s126 Free State 87.8 6.1 6.1 1.98 3.1 8.8

n211 Highveld 55.5 7.43 6.62 3.8 1.7 5.0


Collapsed – ‘normal’ coal

n212 Highveld 78.2 10.53 6.47 5.16 1.3 5.1

n213 Highveld 73.5 8.4 6.6 3.65 1.8 5.9

n214 KwaZulu - Natal 57.66 5.3 5.66 2.89 2.0 6.2

m168 South Rand 165.7 15 5 5.94 0.8 7.4

m169 South Rand 195 17 6 4.88 1.2 8.9

m170 South Rand 205 17 6 5.88 1.0 9.4

s122 South Rand 167.6 15.85 5.49 5.49 1.0 7.6

s66 South Rand 193.2 15.85 5.49 5.49 1.0 8.8

s67 South Rand 184.7 15.85 5.49 5.49 1.0 8.4

s19 Springs-Witbank 36.6 6.1 7.62 4.88 1.6 4.6

s54 Springs-Witbank 62.5 6.1 7.62 2.44 3.1 7.9

n185 Utrecht 101 9 6 3.8 1.6 7.0

n186 Utrecht 100 8.5 6.5 3.3 2.0 7.8

n187 Utrecht 97 9 6.6 3.7 1.8 7.3

n188 Utrecht 51.5 6 6 3.9 1.5 5.2

m148 Witbank 28.5 3.8 5.8 2.7 2.1 4.5

m148a Witbank 34 3.5 6.7 2.7 2.5 7.2

195
ESTIMATING ROCK MASS PROPERTIES

TABLE 11.4 Details of the South African Coal Pillar Database – Collapsed cases continued
Pillar Bord Mining Tributary
Depth
Case Coalfield width width height w/h Load
[m]
[m] [m] [m] [MPa]

m148b Witbank 34 3.5 6.7 2.7 2.5 7.2

m149 Witbank 90 7.5 6 4.8 1.3 7.3

m150 Witbank 57 3.6 5.4 1.35 4.0 8.9

m151 Witbank 62 7.5 6.4 4 1.6 5.3

m162 Witbank 62 7.3 6.2 4 1.6 5.3

m163 Witbank 56 5.1 6.5 3.3 2.0 7.2

m164 Witbank 33 6.4 6.4 4.88 1.3 3.3

m165 Witbank 22 3.5 6.5 1.6 4.1 4.5

m166 Witbank 62 6.1 6.1 4 1.5 6.2

m167 Witbank 62 6.1 6.1 4 1.5 6.2

n171 Witbank 41 6.4 6.4 6.2 1.0 4.1

n172 Witbank 41 6.4 6.4 6.2 1.0 4.1

n173 Witbank 41 6.4 6.4 6.2 1.0 4.1

n174 Witbank 35.5 5.5 5.5 2.2 2.5 3.6

n198 Witbank 32 3.3 6.4 2.3 2.8 6.9


Collapsed – ‘normal’ coal

n199 Witbank 32.5 3.2 6.5 2.1 3.1 7.5

n200 Witbank 43 4.8 6.2 2.8 2.2 5.6

n216 Witbank 86.4 7.5 6.5 4.6 1.4 7.5

n217 Witbank 102 7.6 6.2 4.5 1.4 8.4

s116 Witbank 61 6.1 6.1 4.57 1.3 6.1

s117 Witbank 61 6.1 7.62 3.05 2.5 7.7

s118 Witbank 57.9 6.1 7.62 3.96 1.9 7.3

s119 Witbank 41.1 4.27 6.4 3.05 2.1 6.4

s12 Witbank 25.9 3.66 8.53 3.05 2.8 7.2

s16 Witbank 21.3 3.96 8.23 4.57 1.8 5.0

s17 Witbank 29.6 5.18 7.01 5.49 1.3 4.1

s18 Witbank 27.4 3.66 7.92 2.13 3.7 6.9

s39 Witbank 36.6 4.57 7.62 2.44 3.1 6.5

s40 Witbank 33.5 6.1 6.71 5.49 1.2 3.7

s41 Witbank 30.5 4.57 7.62 3.66 2.1 5.4

s42 Witbank 53.3 5.18 6.4 3.66 1.7 6.7

s55 Witbank 68.6 3.35 5.79 1.52 3.8 12.8

s57 Witbank 88.4 7.16 6.55 4.88 1.3 8.1

s58 Witbank 57.9 5.18 6.4 5.49 1.2 7.2

s64 Witbank 61 4.72 6.86 3.51 2.0 9.2

s9 Witbank 30.5 3.35 6.4 2.59 2.5 6.5

196
11 | Coal Pillar Case Studies

TABLE 11.4 Details of the South African Coal Pillar Database – Collapsed cases continued
Pillar Bord Mining Tributary
Depth
Case Coalfield width width height w/h Load
[m]
[m] [m] [m] [MPa]

s115 Eastern Transvaal 76.2 4.88 6.1 1.37 4.5 9.6

n175 Klip River 70 7.5 5 1.8 2.8 4.9

n176 Klip River 63.5 7.5 5 2.1 2.4 4.4

n177 Klip River 61 6 5 1.9 2.6 5.1

n178 Klip River 61 7.5 5 1.9 2.6 4.2

n179 Klip River 74 10 5 4 1.3 4.2

n197 Klip River 74 7.7 4.8 2 2.4 4.9

m157 Vaal Basin 112 10.55 6.45 2.82 2.3 7.3

m159 Vaal Basin 108 10.55 6.48 3.18 2.0 7.0

n180 Vaal Basin 82 10 5 2.8 1.8 4.6


Collapsed – ‘weak’ coal

n181 Vaal Basin 96 12 6 2.9 2.1 5.4

n182 Vaal Basin 70 12.5 5.5 2.9 1.9 3.6

n183 Vaal Basin 88 11 6 2.9 2.1 5.3

n184 Vaal Basin 112 11.5 5.5 2.9 1.9 6.1

n194 Vaal Basin 96 12 6 6 1.0 5.4

n195 Vaal Basin 82 12 6 3 2.0 4.6

n196 Vaal Basin 104 12 6 3 2.0 5.9

n215 Vaal Basin 128 12.8 5.5 5.5 1.0 6.5

s120 Vaal Basin 128 9.75 5.49 3.66 1.5 7.8

s59 Vaal Basin 57.9 5.18 6.4 3.66 1.7 7.2

s60 Vaal Basin 152.4 12.19 6.1 4.88 1.3 8.6

197
ESTIMATING ROCK MASS PROPERTIES

TABLE 11.5 Details of the South African Coal Pillar Database – Un-collapsed cases
Coalfield Depth [m] Pillar width Bord width Mining w/h Tributary
[m] [m] height [m] Load [MPa]
Highveld 36.2 5.7 6.27 3.8 1.5 4.0
Highveld 39.95 6.89 6.08 3.7 1.9 3.5
Highveld 55.2 8.98 6.96 3.95 2.3 4.3
Highveld 30.35 8.96 6.06 3.85 2.3 2.1
Highveld 41.35 8 5.98 3.6 2.2 3.2
Highveld 57.34 9.07 5.99 3.84 2.4 4.0
KwaZulu-Natal 63.15 8.02 6.11 1.65 4.9 4.9
Witbank 40.86 7.55 7.63 2.92 2.6 4.1
Witbank 30.55 8.07 7.11 3.2 2.5 2.7
Witbank 49.82 7.63 7.67 3.05 2.5 5.0
Witbank 47.07 8.29 6.89 3.37 2.5 3.9
Witbank 66.95 7.77 7.5 3 2.6 6.5
Witbank 25.98 3.11 5.53 1.34 2.3 5.0
Witbank 26.83 2.91 5.09 1.33 2.2 5.1
Witbank 42.61 4.38 5.07 1.61 2.7 5.0
Witbank 23 4 5.68 1.58 2.5 3.4
Witbank 23.76 4.3 5.21 1.5 2.9 2.9
Witbank 23.54 4.24 5.38 1.54 2.8 3.0
Witbank 26.75 4.54 4.97 1.65 2.8 2.9
Witbank 19.4 4.26 5.22 1.4 3.0 2.4
Witbank 20.8 4.24 5.18 1.38 3.1 2.6
Witbank 23.33 4.37 5.08 1.33 3.3 2.7
KwaZulu-Natal 58.89 8.06 5.7 3 2.7 4.3
KwaZulu-Natal 96.2 8.06 5.77 3.1 2.6 7.1
KwaZulu-Natal 33.94 8.17 5.52 2.38 3.4 2.4
Witbank 39.1 6.63 6.36 2.6 2.6 3.8
Witbank 36.5 7.25 6.3 2.68 2.7 3.2
Witbank 24.15 6.1 6.25 2.65 2.3 2.5
Witbank 25.7 6.15 6.17 2.6 2.4 2.6
Witbank 18.58 4.86 6.16 2.78 1.7 2.4
Witbank 21.5 5.09 5.96 2.75 1.9 2.5
Witbank 44.95 5.03 6.01 2.68 1.9 5.4
Witbank 26.51 4.32 6.68 2.65 1.6 4.3
KwaZulu-Natal 97.27 8.63 6.04 1.37 6.3 7.0
KwaZulu-Natal 165.45 11.66 6.32 1.45 8.0 9.8
KwaZulu-Natal 167.68 12.98 6.43 1.45 9.0 9.4
KwaZulu-Natal 183.83 12.99 6.93 1.23 10.6 10.8
KwaZulu-Natal 54 6 6.54 1.2 5.0 5.9
KwaZulu-Natal 112 17 5.5 1.05 16.2 4.9
Witbank 98.15 12.56 5.23 2.65 4.7 4.9
Witbank 64.79 13.03 5.37 3.08 4.2 3.2
Ermelo 80.5 6.58 6.39 1.65 4.0 7.8
Ermelo 54.6 5.67 6.37 1.6 3.5 6.2
Ermelo 25.15 4.39 6.54 1.65 2.7 3.9
Ermelo 35.73 5 6.12 1.73 2.9 4.4

198
11 | Coal Pillar Case Studies

TABLE 11.5 Details of the South African Coal Pillar Database – Un-collapsed cases continued
Coalfield Depth [m] Pillar width Bord width Mining w/h Tributary
[m] [m] height [m] Load [MPa]
Ermelo 55.7 6.85 6.13 1.7 4.0 5.0
Ermelo 52.5 5.66 6.38 1.7 3.3 5.9
Ermelo 35.5 4.6 6.81 2 2.3 5.5
Ermelo 40.33 4.53 6.52 1.83 2.5 6.0
Ermelo 22.62 4.8 6.2 1.65 2.9 3.0
Ermelo 28.75 5 6.08 2 2.5 3.5
KwaZulu-Natal 22.3 6.24 5.74 3.5 1.8 2.1
KwaZulu-Natal 37.7 7.05 5.94 3.06 2.3 3.2
KwaZulu-Natal 36.26 6.23 6.82 2.63 2.4 4.0
KwaZulu-Natal 50.17 5.85 7.04 2.6 2.3 6.1
KwaZulu-Natal 37.08 6.29 6.65 2.78 2.3 3.9
KwaZulu-Natal 61.1 7.08 6.61 4.49 1.6 5.7
KwaZulu-Natal 42 5.38 6.78 3.96 1.4 5.4
KwaZulu-Natal 39.33 8.05 6 2.68 3.0 3.0
KwaZulu-Natal 59.16 7.17 6.57 3.51 2.0 5.4
Free State 64.24 5.59 5.05 1 5.6 5.8
Free State 59.98 6.19 5.86 1.52 4.1 5.7
Free State 17.91 4.31 5.62 1.71 2.5 2.4
Free State 29.56 6.2 6 2.1 3.0 2.9
N.R. 45.72 5.49 5.18 1.83 3.0 4.3
Witbank 76.2 6.71 5.49 2.74 2.4 6.3
N.R. 85.34 8.23 5.49 3.2 2.6 5.9
Witbank 121.92 12.8 5.49 2.29 5.6 6.2
Witbank 97.54 9.75 5.49 3.2 3.0 6.0
Vryheid 108 9 6 1.14 7.9 7.5
Vryheid 108 9 6 1.1 8.2 7.5
Vryheid 150 9 6 1.15 7.8 10.4
Vryheid 208 17 6 1.2 14.2 9.5
Vryheid 208 17 6 1.1 15.5 9.5
Vryheid 166 17 6 1.48 11.5 7.6
Vryheid 174 17 6 1.17 14.5 8.0
Witbank 64.47 9 6 1.89 4.8 4.5
Witbank 50 10 6 3.8 2.6 3.2
Witbank 34.6 12.21 5.6 3.4 3.6 1.8
Witbank 77.65 12 5.84 3.8 3.2 4.3
Witbank 55.75 10.28 6 2.8 3.7 3.5
Witbank 55.75 11 6 3 3.7 3.3
Witbank 79.92 9 6 2.9 3.1 5.6
Witbank 34.15 8 6 3 2.7 2.6
Witbank 36.53 7 6 3 2.3 3.1
Witbank 45.72 4.27 5.49 1.19 3.6 6.0
Witbank 68.58 5.18 5.49 1.19 4.4 7.3
Witbank 91.44 6.1 5.49 1.19 5.1 8.3
Witbank 76.2 7.62 6.1 1.37 5.6 6.2
Witbank 30.48 5.49 6.71 2.59 2.1 3.8

199
ESTIMATING ROCK MASS PROPERTIES

TABLE 11.5 Details of the South African Coal Pillar Database – Un-collapsed cases continued
Coalfield Depth [m] Pillar width Bord width Mining w/h Tributary
[m] [m] height [m] Load [MPa]
Witbank 63.64 14 6 3 4.7 3.2
Witbank 70.79 12 6 2.2 5.5 4.0
Witbank 118.62 15 6 3.5 4.3 5.8
Witbank 108.75 14 6 3 4.7 5.5
Witbank 53.86 7 6 5 1.4 4.6
Witbank 98.72 13 6 4 3.3 5.3
Highveld 170 22 6 2.6 8.5 6.9
Highveld 154 21.37 6.26 2.72 7.9 6.4
Highveld 171.7 21.19 6.81 2.9 7.3 7.5
Highveld 151 23.85 6.15 3.34 7.1 6.0
Highveld 137 21.82 5.84 2.84 7.7 5.5
Highveld 106 15.84 6.35 3.16 5.0 5.2
Vereeniging 91.44 6.1 6.1 1.52 4.0 9.1
Vereeniging 60.96 9.75 5.49 2.29 4.3 3.7
Vereeniging 114 12.8 5.49 1.98 6.5 5.8
Vereeniging 146.3 17.37 5.49 2.13 8.2 6.3
Vereeniging 91.44 12.8 5.49 2.07 6.2 4.7
Vereeniging 114 17.37 5.49 1.98 8.8 4.9
Vereeniging 106.68 9.14 6.1 1.68 5.4 7.4
Vereeniging 140.21 12.8 5.49 2.44 5.2 7.2
Vereeniging 93.88 9.75 5.49 1.98 4.9 5.7
Witbank 39.62 6.1 6.1 2.59 2.4 4.0
Witbank 91.44 6.1 6.1 1.55 3.9 9.1
Klip River 223.38 28 5 3.5 8.0 7.8
Klip River 173.22 25 5 4.2 6.0 6.2
Klip River 215.05 35 5 3 11.7 7.0
Klip River 254.45 35 5 4 8.8 8.3
Klip River 215.05 25 5 3 8.3 7.7
Witbank 62.48 10 6 2.2 4.5 4.0
Witbank 62.48 15 6 2.1 7.1 3.1
Witbank 62.48 9 6 2.25 4.0 4.3
Witbank 70 15 6 3 5.0 3.4
Witbank 122.17 12 6 2.8 4.3 6.9
Witbank 122 15.4 6.6 2.5 6.2 6.2
Witbank 122 15.4 6.6 2.5 6.2 6.2
Highveld 94 10.22 6.78 3 3.4 6.5
Highveld 94 10.2 6.8 2.8 3.6 6.5
Witbank 45.72 6.1 6.1 2.9 2.1 4.6
Witbank 60.96 7.62 6.1 2.9 2.6 4.9
Witbank 76.2 9.14 6.1 2.9 3.2 5.3
Witbank 91.44 10.67 6.1 2.9 3.7 5.6
Witbank 40.07 8 6 3.2 2.5 3.1
Witbank 40.07 7.5 6 3.2 2.3 3.2
Witbank 56.11 7.5 6 3.5 2.1 4.5
Witbank 111.91 13.4 6 3 4.5 5.9

200
11 | Coal Pillar Case Studies

TABLE 11.5 Details of the South African Coal Pillar Database – Un-collapsed cases continued
Coalfield Depth [m] Pillar width Bord width Mining w/h Tributary
[m] [m] height [m] Load [MPa]
Witbank 121.02 13 6 3.5 3.7 6.5
Witbank 143.26 12.8 5.49 1.68 7.6 7.3
Eastern Tvl 140.72 15 6 2.7 5.6 6.9
Eastern Tvl 83.43 12 6 2.8 4.3 4.7
Eastern Tvl 101.21 12 6 2.3 5.2 5.7
Eastern Tvl 114.35 12 6 2.3 5.2 6.4
Eastern Tvl 119.41 15 6 2.3 6.5 5.9
Eastern Tvl 114.3 12 6 2.6 4.6 6.4
Eastern Tvl 104.76 12 6 4.8 2.5 5.9
Eastern Tvl 130.75 15 6 2 7.5 6.4
Witbank 44.55 12 6 2.5 4.8 2.5
Witbank 44.55 12 6 3 4.0 2.5
Witbank 40.3 8 6 2.1 3.8 3.1
Witbank 37.36 10 6 2.8 3.6 2.4
Witbank 70.76 14 6 2.6 5.4 3.6
Witbank 44.55 6 6 3.3 1.8 4.5
Witbank 87.53 12 6 3.6 3.3 4.9
Witbank 90.22 8.53 6.71 2.9 2.9 7.2
Witbank 115.82 7.62 6.1 1.83 4.2 9.4
Witbank 84.84 7 6 2.26 3.1 7.3
Witbank 84.84 7 6 2.26 3.1 7.3
Witbank 84.84 7 6 2.26 3.1 7.3
Witbank 61.5 9.47 6 1.98 4.8 4.1
Witbank 61.5 9.26 6 1.98 4.7 4.2
Witbank 58.71 11 6 2.3 4.8 3.5
Witbank 58.71 7 6 2 3.5 5.1
Witbank 108.63 12.5 6 3.38 3.7 5.9
Witbank 108.63 18.5 6 3.65 5.1 4.8
Vryheid 94.28 10 5 1.4 7.1 5.3
Vryheid 94.28 10 5 1.7 5.9 5.3
Witbank 51.82 9.14 6.1 1.68 5.4 3.6
Witbank 64.01 7.62 7.62 3.05 2.5 6.4
Witbank 48.77 9.14 6.1 4.27 2.1 3.4
Witbank 108.51 8.53 6.71 2.29 3.7 8.7
Highveld 46.1 7.1 6.9 3.8 1.9 4.5
Highveld 50 9.3 5.7 3.65 2.5 3.3
Highveld 58.3 8.47 6 3.8 2.2 4.3
Highveld 59.3 11 6 3.7 3.0 3.5
Utrecht 196 15 6 2.6 5.8 9.6
Highveld 64.92 20.57 6 3.5 5.9 2.7
Highveld 55 11.8 6.2 3.41 3.5 3.2
Highveld 65 10.98 6.02 3.2 3.4 3.9
Highveld 51 8.89 6.11 3.27 2.7 3.6
Highveld 68 11.66 6.34 3.4 3.4 4.1
Highveld 46 8.6 6.4 3.55 2.4 3.5

201
ESTIMATING ROCK MASS PROPERTIES

TABLE 11.5 Details of the South African Coal Pillar Database – Un-collapsed cases continued
Coalfield Depth [m] Pillar width Bord width Mining w/h Tributary
[m] [m] height [m] Load [MPa]
Highveld 74 11.31 5.69 3.41 3.3 4.2
Highveld 68 10.87 6.13 3.37 3.2 4.2
Highveld 68 12.9 7.1 4.06 3.2 4.1
Highveld 54 8.92 6.08 3.58 2.5 3.8
Highveld 70 10.68 6.32 3.25 3.3 4.4
Highveld 49 8.57 6.43 3.69 2.3 3.8
Highveld 51 8.65 6.35 3.64 2.4 3.8
Highveld 72 12.89 7.11 4.15 3.1 4.3
Highveld 77 10.04 5.96 2.94 3.4 4.9
Highveld 57 9.35 6.65 3.65 2.6 4.2
Highveld 75 10.7 6.3 3.41 3.1 4.7
Highveld 78 10.7 6.3 3.35 3.2 4.9
Highveld 54 8.52 6.48 3.69 2.3 4.2
Highveld 56 8.59 6.41 3.63 2.4 4.3
Highveld 64.92 9 6 3.6 2.5 4.5
Highveld 73 10.63 6.37 3.85 2.8 4.7
Highveld 81 13.01 6.99 4.31 3.0 4.8
Highveld 64.92 9 6 3.72 2.4 4.5
Highveld 75 10.03 6.97 3.07 3.3 5.4
Highveld 84 12.97 7.03 4.31 3.0 5.0
Highveld 79 13.58 6.42 5.89 2.3 4.3
Highveld 74 9.9 7.1 3.33 3.0 5.5
Highveld 71 12.32 6.68 5.84 2.1 4.2
Highveld 73 10.21 6.79 3.93 2.6 5.1
Highveld 83 13.66 6.34 6.13 2.2 4.4
Highveld 75 11.03 5.97 5.5 2.0 4.5
Highveld 77 10.22 6.78 4.05 2.5 5.3
Highveld 71 10.7 6.3 5.5 1.9 4.5
Highveld 76 10.8 6.2 5.5 2.0 4.7
Highveld 66 10.92 7.08 6.45 1.7 4.5
Highveld 76 8.85 7.15 3.48 2.5 6.2
Highveld 77 9.81 7.19 4.86 2.0 5.8
Highveld 93.8 14 6 3.6 3.9 4.8
Highveld 93.8 14.92 6 3 5.0 4.6
Highveld 93.8 14.7 6 3.2 4.6 4.6
Highveld 109 17.5 6.48 3.67 4.8 5.1
Highveld 106 16.47 7.5 3.5 4.7 5.6
Highveld 92.28 17.07 6.93 3.87 4.4 4.6
Highveld 122 21 6.71 3.29 6.4 5.3
Highveld 122 21.2 6.71 3.29 6.4 5.3
N.R. 97.54 9.14 6.1 3.2 2.9 6.8
N.R. 106.68 15.24 6.1 4.88 3.1 5.2
N.R. 45.72 4.88 6.1 1.65 3.0 5.8
N.R. 60.96 4.88 6.1 1.55 3.1 7.7
N.R. 60.96 7.01 6.71 2.59 2.7 5.8

202
11 | Coal Pillar Case Studies

TABLE 11.5 Details of the South African Coal Pillar Database – Un-collapsed cases continued
Coalfield Depth [m] Pillar width Bord width Mining w/h Tributary
[m] [m] height [m] Load [MPa]
N.R. 30.48 2.74 5.49 1.55 1.8 6.9
N.R. 85.34 7.62 6.1 3.2 2.4 6.9
N.R. 167.64 14.33 3.69 1.98 7.2 6.6
N.R. 108.51 7.62 6.71 2.29 3.3 9.6
N.R. 42.67 4.27 5.49 2.59 1.6 5.6
N.R. 140.21 17.37 5.49 2.44 7.1 6.1
N.R. 160.93 14.94 6.31 2.96 5.0 8.1
N.R. 180.44 15.24 5.97 2.99 5.1 8.7
N.R. 163.98 15 6.22 2.8 5.4 8.2
N.R. 146 16.46 5.49 2.41 6.8 6.5
N.R. 198.12 17.16 5.7 2.83 6.1 8.8
N.R. 182.88 15.85 5.49 4.88 3.2 8.3
N.R. 36.58 7.32 4.88 1.83 4.0 2.5
N.R. 91.44 7.62 6.1 1.68 4.5 7.4
N.R. 60.96 9.14 6.1 4.88 1.9 4.2
N.R. 76.2 10.67 6.1 4.88 2.2 4.7
N.R. 91.44 12.19 6.1 4.88 2.5 5.1
Witbank 45.72 6.1 6.1 3.05 2.0 4.6
Highveld 178.49 19 6 2.3 8.3 7.7
Highveld 195.15 20 6 2.1 9.5 8.2
Highveld 203.42 19 6 1.7 11.2 8.8
Highveld 203.42 19 6 1.7 11.2 8.8
Witbank 42.67 5.49 5.49 2.59 2.1 4.3
N.R. 91.44 6.1 6.1 1.68 3.6 9.1
Witbank 48.77 7.32 6.4 2.44 3.0 4.3
Witbank 83.92 11.47 6 3 3.8 4.9
Witbank 87.6 12 6 2.8 4.3 4.9
Witbank 45.72 6.1 6.1 3.96 1.5 4.6
Witbank 60.96 7.62 6.1 3.96 1.9 4.9
Witbank 45.72 6.1 6.1 1.68 3.6 4.6
Witbank 60.96 7.62 6.1 1.68 4.5 4.9
Witbank 76.2 9.14 6.1 3.96 2.3 5.3
Witbank 91.44 12.19 6.1 3.96 3.1 5.1
Witbank 76.2 9.14 6.1 1.68 5.4 5.3
Witbank 33.53 6.71 5.49 3.96 1.7 2.8
Witbank 115.82 9.14 6.1 1.83 5.0 8.1
Highveld 169 21.81 6.19 2.58 8.5 7.0
Vereeniging 70 12 6 2.6 4.6 3.9
Vereeniging 104 11 6 3 3.7 6.2
Vereeniging 110.5 12 5 2.8 4.3 5.5
Vereeniging 110.5 12 5 3 4.0 5.5
Vereeniging 70 20.57 5 2.6 7.9 2.7
Vereeniging 70 15.44 5 2.6 5.9 3.1
Vereeniging 93 10 6 2.9 3.4 6.0
Vereeniging 84 11 5.89 2.77 4.0 5.0

203
ESTIMATING ROCK MASS PROPERTIES

TABLE 11.5 Details of the South African Coal Pillar Database – Un-collapsed cases continued
Coalfield Depth [m] Pillar width Bord width Mining w/h Tributary
[m] [m] height [m] Load [MPa]
Vereeniging 84 11.5 5.89 2.77 4.2 4.8
Witbank 47.24 6.1 7.62 5.18 1.2 6.0
Witbank 64.01 6.1 6.25 3.44 1.8 6.6
Witbank 91.44 10.67 6.1 4.27 2.5 5.6
Witbank 38.1 4.27 5.49 1.52 2.8 5.0
Witbank 106.68 12.19 6.1 4.27 2.9 6.0
Witbank 88.39 9.14 6.1 3.05 3.0 6.1
Klip River 84.65 6 6 1.75 3.4 8.5
Witbank 85.34 7.62 6.1 3.05 2.5 6.9
Witbank 108.51 8.84 6.4 2.29 3.9 8.1
Soutpansberg 79.4 24 6 3 8.0 3.1
Soutpansberg 138.89 17.78 6 3 5.9 6.2
Soutpansberg 138.89 31.06 6 3 10.4 4.9
Soutpansberg 79.4 19.47 6 3 6.5 3.4
Witbank 60.96 7.01 6.71 3.35 2.1 5.8
Witbank 30.48 6.1 6.1 3.2 1.9 3.0
Witbank 44.2 6.1 6.1 4.27 1.4 4.4
Witbank 42.8 7 6 2.2 3.2 3.7
Witbank 67.8 11 6 4 2.8 4.0
Witbank 67.8 11.73 6 4 2.9 3.9
Witbank 36.3 8 6 3.7 2.2 2.8
Highveld 51.8 10.98 6.02 2.88 3.8 3.1
Highveld 124.9 18.27 5.73 2.67 6.8 5.4
Highveld 110 17.82 6.18 3.46 5.2 5.0
Highveld 146.2 19.89 6.29 2.91 6.8 6.3
Witbank 68.9 12.9 6 3 4.3 3.7
Witbank 34.5 5.5 6.5 5.5 1.0 4.1
Witbank 33.8 8 6 2.4 3.3 2.6
Witbank 52.8 8 6 3 2.7 4.0
Witbank 38.37 8.1 6.4 5.9 1.4 3.1
Witbank 52.8 11.3 6 3 3.8 3.1
Witbank 45.72 6.1 6.1 4.57 1.3 4.6
Witbank 219.46 21.73 5.58 3.17 6.9 8.7
Witbank 76.2 7.62 6.1 4.57 1.7 6.2
Witbank 106.68 6.71 5.49 1.19 5.6 8.8
Free State 182.88 16.92 5.94 2.44 6.9 8.3
Free State 91.44 12.19 6.1 1.52 8.0 5.1
Free State 35.05 9.75 5.49 2.44 4.0 2.1
Free State 50.6 6.1 6.1 1.37 4.5 5.1
Free State 76.2 9.14 6.1 1.37 6.7 5.3
Free State 35.05 7.32 4.88 2.59 2.8 2.4
Free State 42 5 10 1.8 2.8 9.5
Free State 36 5 7 2.4 2.1 5.2
Free State 38 5 7 1.8 2.8 5.5
Free State 22 6 6 1.5 4.0 2.2

204
11 | Coal Pillar Case Studies

TABLE 11.5 Details of the South African Coal Pillar Database – Un-collapsed cases continued
Coalfield Depth [m] Pillar width Bord width Mining w/h Tributary
[m] [m] height [m] Load [MPa]
Witbank 36.58 6.1 6.1 3.66 1.7 3.7
Witbank 51.82 7.62 6.1 3.66 2.1 4.2
Witbank 137.16 12.8 5.49 3.66 3.5 7.0
Witbank 91.44 9.14 6.1 4.57 2.0 6.4
Witbank 45.72 7.62 6.1 4.88 1.6 3.7
Witbank 19.81 7.62 6.1 3.2 2.4 1.6
Utrecht 29.77 14 5 3 4.7 1.4
Utrecht 58.53 11.42 6 3.5 3.3 3.4
Utrecht 58.53 14 6 3.5 4.0 3.0
Utrecht 42.59 15 6 3.7 4.1 2.1
Utrecht 13.22 8 6 2.6 3.1 1.0
Utrecht 65.78 10 5 2.2 4.5 3.7
Utrecht 54.65 10 5 2.2 4.5 3.1
Utrecht 54.99 9.47 5 3 3.2 3.2
Witbank 45.72 6.55 6.4 3.51 1.9 4.5
Witbank 82 12 6 3 4.0 4.6
Witbank 70.1 12 6 3.35 3.6 3.9
Witbank 21.34 6.1 6.71 2.74 2.2 2.4
Witbank 30.48 6.4 6.4 3.66 1.7 3.0
Witbank 41.15 6.4 6.4 1.98 3.2 4.1
Zululand 78.54 14 6 2.7 5.2 4.0
Zululand 78.54 14 6 2.4 5.8

205
ESTIMATING ROCK MASS PROPERTIES

11.2.3 USA (1999); Mark (2006) summarised three modes of coal


The USA National Institute of Occupational Safety pillar behaviour noting that predicting pillar strength
and Health (NIOSH) collated 645 case histories of becomes more difficult as the ratio of pillar width to
retreat mining workings. It makes the data available as height (w/h) increases.
part of its Analysis of Retreat Mining Pillar Stability • Sudden, massive collapse for slender pillars; w/h < 4
(ARMPS) software (14). That data has been filtered to • Squeezing, non-violent failure; 4 < w/h <10
include only those 170 cases of development workings, • Squat ‘indestructible’ pillars which have w/h > 10.
to equate the USA database to the bord & pillar There are six definite outliers in the USA database
workings of the Australian and South African cases. of un-collapsed pillars. These pillars can be seen
The data is summarised in Table 11.6 and Figure within the annotated ellipse in Figure 11.3 with w/h
11.3. It is detailed in Table 11.7 for the 35 collapsed ≈ 3 and at between 425 and 600 m depth. The six
cases and in Table 11.8 for the 135 un-collapsed cases. cases come from two mines in Utah which work
NIOSH defines ‘massive collapses’ as pillar failures highly rectangular pillars, the dimensions of which
that take place rapidly and involve large areas. It is are 9 x 34 m and 11 x 31 m (as highlighted in Table
worth noting that all the massive collapses involved 11.8). In fact, the ratios between the pillar dimensions
geometries with w/h < 3. This ratio was used to define (3.77 and 2.70, respectively) are the largest of the
‘slender pillars’ by Mark (1999) but the ratio was un-collapsed database.
subsequently revised to w/h < 4 (Mark 2006). Mark Galvin et al. (1999) recognised the influence of

TABLE 11.6 Summary of the USA coal pillar database


Case Number of Depth [m] w/h
cases

Squeeze 23 76 – 488 1.0 – 8.0

Collapsed Massive collapse 11 53 – 168 1.5 – 2.5

Local burst 1 472 5.9

Un-collapsed 135 64– 594 3.0 – 16.7

w/h
0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0
0

50

100

150

200

250
Depth [m]

300

350

400

450

500
Uncollapsed
Collapsed - squeeze
550
Collapsed - massive
Collapsed - Local burst
600

FIGURE 11.3 The database of USA development workings coal pillars collated by NIOSH

(14)
ARMPS(2010) v.6.2.02 August 2013 downloaded from http://www.cdc.gov/niosh/mining/works/coversheet1813.html on 6th November 2014

206
11 | Coal Pillar Case Studies

rectangular pillars and recommended calculating an If this approach was adopted for the six outlier cases,
effective pillar width as the harmonic mean according then their w/h would increase to 4.6 to 6.0. 
to Equation 5.10 where w1 ≤ w2 .

EQUATION 11.3

TABLE 11.7 Details of the USA coal pillar database – Collapsed cases (the last 12 cases are those defined
as massive collapses or rock burst and are shown in italics)
Case Depth Pillar Width Pillar w/h Bord
[m] m] Height Width
State Mine (minimum width) [m] [m]
KY 20 442 15.2 x 9.1 9.1 1.8 5.0 6.1
KY 20 396 15.2 x 9.1 9.1 1.7 5.5 6.1
KY 42 320 15.5 x 11 11.0 1.8 6.0 5.8
KY 60 396 18.6 x 11 11.0 1.8 6.2 5.8
KY 74 326 22.6 x 16.5 16.5 3.7 4.5 5.5
KY 75 122 9.1 x 9.1 9.1 2.0 4.6 6.1
KY 91 244 15.2 x 12.2 12.2 2.3 5.3 6.1
KY 128 396 22.9 x 22.9 22.9 3.4 6.8 6.1
KY 128 488 29 x 24.4 24.4 3.2 7.6 6.1
KY 137 244 15.2 x 15.2 15.2 4.0 3.8 6.1
OH 92 158 9.8 x 7.7 7.7 1.8 4.2 5.5
OH 92 198 9.8 x 7.7 7.7 1.8 4.2 5.5
UT 59 305 18.3 x 12.2 12.2 2.0 6.2 6.1
VA 16 488 24.4 x 16.8 16.8 4.0 4.2 6.1
WV 2 91 3 x 12.2 3.0 2.4 1.3 6.1
WV 67 351 18.3 x 15.2 15.2 1.9 8.0 6.1
WV 79 130 12.2 x 3 3.0 2.4 1.3 6.1
Collapsed

WV 79 122 12.2 x 3 3.0 2.5 1.2 6.1


WV 81 76 12.2 x 3 3.0 2.1 1.5 6.1
WV 82 122 12.2 x 3 3.0 3.0 1.0 6.1
WV 133 122 12.2 x 3 3.0 1.5 2.0 6.1
WV 133 114 12.2 x 3 3.0 1.4 2.2 6.1
WV 134 76 13.7 x 3.8 3.8 2.1 1.8 6.1
CO 111 91 4.1 x 19.4 4.1 2.1 1.9 11.1
UT 11 168 12.2 x 12.2 12.2 5.3 2.3 6.1
WV 23 73 12.2 x 3 3.0 3.0 1.0 6.1
WV 23 73 18.3 x 3 3.0 3.0 1.0 6.1
WV 23 85 6.1 x 12.2 6.1 3.0 2.0 6.1
WV 23 75 3 x 12.2 3.0 3.0 1.0 6.1
WV 36 69 6.1 x 6.1 6.1 3.4 1.8 6.1
WV 57 53 12.2 x 3 3.0 3.0 1.0 6.1
WV 76 91 6.1 x 6.1 6.1 2.4 2.5 6.1
WV 82 84 12.2 x 3 3.0 2.9 1.1 6.1
WV 82 84 12.2 x 3 3.0 2.9 1.1 6.1
CO 116 472 18.3 x 15.2 15.2 2.6 5.9 6.1

207
ESTIMATING ROCK MASS PROPERTIES

TABLE 11.8 Details of the USA coal pillar database – Un-collapsed cases
Case Depth Pillar Width Pillar w/h Bord
[m] m] Height Width
State Mine (minimum width) [m] [m]
KY 21 269 18.3 x 12.2 12.2 1.8 6.7 6.1
KY 21 329 18.3 x 12.2 12.2 1.7 7.1 6.1
KY 21 271 18.3 x 18.3 18.3 1.4 12.8 6.1
KY 21 361 18.3 x 21.3 18.3 1.5 12.5 6.1
KY 21 354 18.3 x 21.3 18.3 1.6 11.1 6.1
KY 21 344 18.3 x 15.2 15.2 1.8 8.5 6.1
KY 21 341 12.2 x 15.2 12.2 1.5 8.3 6.1
KY 21 305 18.3 x 12.2 12.2 1.5 8.3 6.1
KY 22 224 18.3 x 18.3 18.3 1.1 16.7 6.1
KY 22 225 18.3 x 18.3 18.3 1.2 15.8 6.1
KY 22 293 18.3 x 18.3 18.3 1.3 13.6 6.1
KY 22 193 14.9 x 19.8 14.9 1.2 12.7 6.1
KY 25 234 19.8 x 15.2 15.2 2.0 7.7 6.1
KY 25 156 15.2 x 15.2 15.2 2.0 7.8 6.1
KY 25 88 14.9 x 15.2 14.9 1.9 7.8 6.1
KY 25 117 18.3 x 19.8 18.3 1.9 9.5 6.1
KY 25 234 18.3 x 21.3 18.3 2.0 9.2 6.1
KY 25 156 18.3 x 36.6 18.3 2.0 9.4 6.1
KY 25 187 19.2 x 27.4 19.2 2.7 7.1 6.1
KY 28 305 21.3 x 21.6 21.3 1.6 13.3 6.1
KY 30 238 14.9 x 15.2 14.9 1.6 9.2 6.1
KY 30 184 19.2 x 18.3 18.3 1.5 12.5 6.1
Un-collapsed

KY 32 546 18.3 x 16.8 16.8 1.5 11.0 6.1


KY 32 424 18.3 x 15.2 15.2 1.6 9.6 6.1
KY 58 427 24.4 x 27.4 24.4 1.6 15.1 6.1
KY 61 189 12.5 x 15.2 12.5 2.2 5.6 6.1
KY 61 436 18.3 x 18.3 18.3 2.2 8.2 6.1
KY 61 433 18.3 x 24.4 18.3 2.2 8.2 6.1
KY 61 490 18.3 x 24.4 18.3 2.2 8.2 6.1
KY 77 537 18.3 x 18.3 18.3 1.7 10.9 6.1
KY 85 274 12.2 x 9.1 9.1 1.6 5.8 6.1
KY 100 274 12.2 x 9.1 9.1 1.3 6.9 6.1
KY 100 274 12.2 x 10.7 10.7 1.3 8.3 6.1
KY 100 236 9.1 x 9.1 9.1 1.3 7.0 6.1
KY 100 290 9.1 x 6.1 6.1 1.3 4.6 6.1
KY 100 274 9.1 x 9.1 9.1 1.6 5.8 6.1
KY 100 290 9.1 x 9.1 9.1 1.6 5.7 6.1
KY 100 274 12.2 x 10.7 10.7 1.4 7.8 6.1
KY 100 320 12.2 x 12.2 12.2 1.4 8.4 6.1
KY 100 274 12.2 x 12.2 12.2 1.6 7.6 6.1
KY 100 297 12.2 x 9.1 9.1 1.4 6.3 6.1
KY 100 305 9.1 x 9.1 9.1 1.4 6.3 6.1
KY 114 152 18.3 x 18.3 18.3 1.4 13.3 6.1
KY 114 217 27.4 x 16.8 16.8 1.4 12.2 6.1
KY 114 217 19.2 x 27.4 19.2 1.4 14.0 6.1
KY 131 335 18.6 x 9.4 9.4 1.7 5.6 5.8

208
11 | Coal Pillar Case Studies

TABLE 11.8 Details of the USA coal pillar database – Un-collapsed cases continued
Case Depth Pillar Width Pillar w/h Bord
[m] m] Height Width
State Mine (minimum width) [m] [m]
KY 131 366 12.5 x 12.5 12.5 1.7 7.5 5.8
OH 92 181 9.8 x 7.7 7.7 1.8 4.2 5.5
OH 92 164 9.8 x 7.7 7.7 1.8 4.2 5.5
OH 92 204 11.3 x 9 9.0 1.8 4.9 5.5
OH 92 161 9.8 x 7.7 7.7 1.8 4.2 5.5
OH 92 196 11.3 x 11.3 11.3 1.8 6.2 5.5
OH 92 158 9.8 x 7.7 7.7 1.8 4.2 5.5
OH 92 154 9.8 x 7.7 7.7 1.8 4.2 5.5
OH 92 153 9.8 x 7.7 7.7 1.8 4.2 5.5
OH 92 157 9.8 x 7.7 7.7 1.8 4.2 5.5
OH 92 179 15.8 x 12.8 12.8 1.8 7.0 5.5
OH 92 162 15.8 x 12.8 12.8 2.4 5.3 5.5
OH 92 183 15.8 x 12.8 12.8 1.8 7.0 5.5
PA 99 64 5.2 x 5.2 5.2 1.3 4.0 5.5
PA 113 76 12.5 x 12.5 12.5 1.5 8.2 5.8
PA 120 260 24.4 x 21.3 21.3 2.1 10.0 5.2
UT 3 585 11.3 x 30.5 11.3 2.9 3.9 6.1
UT 4 207 20.4 x 18.3 18.3 2.1 8.6 6.1
UT 4 259 20.4 x 18.3 18.3 2.1 8.6 6.1
UT 5 137 12.5 x 12.2 12.2 1.4 8.9 6.1
UT 5 244 18.3 x 21.3 18.3 1.4 13.3 6.1
UT 5 244 18.3 x 18.3 18.3 1.4 13.3 6.1
Un-collapsed

UT 5 114 18.3 x 18.3 18.3 1.4 13.3 6.1


UT 5 183 20.4 x 18.3 18.3 1.4 13.3 6.1
UT 5 259 20.4 x 18.3 18.3 1.4 13.3 6.1
UT 35 579 34.4 x 9.1 9.1 3.0 3.0 6.1
UT 35 594 34.4 x 9.1 9.1 3.0 3.0 6.1
UT 35 503 34.4 x 9.1 9.1 3.0 3.0 6.1
UT 35 427 34.4 x 9.1 9.1 3.0 3.0 6.1
UT 35 549 34.4 x 9.1 9.1 3.0 3.0 6.1
UT 121 305 9.1 x 24.4 9.1 2.7 3.3 6.1
VA 8 518 18.3 x 18.3 18.3 1.7 10.9 6.1
VA 52 213 14.3 x 13.7 13.7 1.1 12.9 6.1
VA 52 168 16.8 x 24.4 16.8 1.1 15.7 6.1
VA 52 305 14.3 x 13.7 13.7 1.1 12.9 6.1
VA 52 183 14.3 x 30.5 14.3 1.1 13.4 6.1
VA 52 244 14.3 x 13.7 13.7 1.1 12.9 6.1
VA 64 282 18.3 x 18.3 18.3 1.3 14.1 6.1
VA 64 244 18.3 x 18.3 18.3 1.3 14.1 6.1
VA 64 314 18.3 x 18.3 18.3 1.3 14.1 6.1
VA 64 314 18.3 x 18.3 18.3 1.3 14.1 6.1
VA 64 274 18.3 x 15.2 15.2 1.3 11.8 6.1
VA 64 335 24.4 x 18.3 18.3 1.3 14.1 6.1
VA 64 335 24.4 x 18.3 18.3 1.3 14.1 6.1
VA 64 282 18.3 x 18.3 18.3 1.3 14.1 6.1
VA 64 174 24.4 x 15.2 15.2 1.3 11.8 6.1

209
ESTIMATING ROCK MASS PROPERTIES

TABLE 11.8 Details of the USA coal pillar database – Un-collapsed cases continued
Case Depth Pillar Width Pillar w/h Bord
[m] m] Height Width
State Mine (minimum width) [m] [m]
VA 64 296 19.2 x 15.2 15.2 1.3 11.8 6.1
VA 64 280 18.3 x 18.3 18.3 1.3 14.1 6.1
VA 64 238 18.3 x 18.3 18.3 1.3 14.1 6.1
VA 64 305 14.3 x 18.3 14.3 1.3 11.1 6.1
VA 64 305 24.4 x 18.3 18.3 1.3 14.1 6.1
VA 64 244 24.4 x 18.3 18.3 1.3 14.1 6.1
VA 88 396 24.4 x 21.3 21.3 1.5 14.0 6.1
VA 88 411 17.4 x 15.2 15.2 1.5 10.0 6.1
VA 112 198 14.9 x 12.2 12.2 1.4 8.4 6.1
VA 112 209 14.9 x 12.2 12.2 1.4 8.4 6.1
VA 112 233 19.2 x 12.2 12.2 1.4 8.4 6.1
VA 112 197 19.2 x 12.2 12.2 1.4 8.4 6.1
VA 112 122 19.2 x 12.2 12.2 1.4 8.4 6.1
VA 141 145 21.6 x 15.2 15.2 1.1 14.3 6.1
VA 141 152 21.6 x 15.2 15.2 1.1 14.3 6.1
VA 141 163 12.5 x 9.1 9.1 1.1 8.6 6.1
VA 141 158 21.6 x 9.1 9.1 1.1 8.6 6.1
VA 141 183 19.2 x 9.1 9.1 1.1 8.6 6.1
VA 141 160 19.2 x 9.1 9.1 1.1 8.6 6.1
WV 6 210 21.6 x 21.3 21.3 1.6 13.2 6.1
WV 6 164 24.4 x 15.2 15.2 1.6 9.4 6.1
WV 33 198 18.3 x 21.3 18.3 3.5 5.2 6.1
WV 33 175 21.6 x 21.3 21.3 3.9 5.4 6.1
WV 33 171 18.3 x 18.3 18.3 3.6 5.1 6.1
WV 40 333 18.3 x 27.4 18.3 2.0 9.0 6.1
WV 40 335 21.6 x 18.3 18.3 1.9 9.7 6.1
WV 40 335 24.4 x 24.4 24.4 1.9 12.7 6.1
WV 40 326 18.3 x 24.4 18.3 2.0 9.2 6.1
WV 66 305 18.3 x 15.2 15.2 1.8 8.3 6.1
WV 70 283 16.5 x 18.3 16.5 2.0 8.2 6.1
WV 70 240 14.9 x 24.4 14.9 2.0 7.4 6.1
WV 83 219 14.9 x 15.2 14.9 1.6 9.2 6.1
WV 118 122 18.3 x 18.3 18.3 2.0 9.0 6.1
WV 118 114 18.3 x 12.2 12.2 1.6 7.4 6.1
WV 118 114 18.3 x 12.2 12.2 1.6 7.4 6.1
WV 144 186 12.5 x 18.3 12.5 1.0 12.6 6.1
WV 144 219 12.5 x 18.3 12.5 1.0 12.6 6.1
WV 144 226 12.5 x 18.3 12.5 1.0 12.6 6.1
WV 144 146 12.5 x 18.3 12.5 1.0 12.6 6.1
WV 144 215 12.5 x 18.3 12.5 1.0 12.6 6.1
WV 144 223 12.5 x 18.3 12.5 1.0 12.6 6.1
WV 144 192 12.5 x 18.3 12.5 1.0 12.6 6.1
WV 144 143 12.5 x 18.3 12.5 1.0 12.6 6.1

210
11 | Coal Pillar Case Studies

11.2.4 India The cases at 450 m depth all come from the Jitpur
Sheorey et al. (1987) collated the details of 23 shaft XIV seam which has weaker coal (Sheorey et al.
collapsed and 20 un-collapsed cases from Indian coal 1987). Further, numerical analyses reported in Sheorey
mines ranging from 23 to 450 m depth. The data is and Singh (1982) show that in this seam ‘virtually
summarised in Table 11.9 and Figure 11.4. It is detailed the whole shaft pillar area has been in a state of post-
in Table 11.10. failure squeezing’. It is inferred from this that the ‘un-
collapsed’ pillars from the Jitpur mine are yielding.

TABLE 11.9 Summary of the Indian coal pillar database


Case Number of Depth [m] w/h
cases

Collapsed Stable 23 23 – 450 0.6 – 6.7

Stable 15 30 – 270 1.4 – 8.3


Un-collapsed
Yielding 5 450 6.4 – 7.6

w/h
0.00 2.00 4.00 6.00 8.00 10.00 12.00 14.00 16.00 18.00
0

50

100

150

200

250
Depth [m]

300

350

400

450

500 Un-collapsed
Collapsed
550
Yielding

600

FIGURE 11.4 The database of Indian coal pillars

211
ESTIMATING ROCK MASS PROPERTIES

TABLE 11.10 Details of the Indian coal pillar database


Case Depth Pillar Width Pillar w/h Bord
[m] m] Height Width
Mine Seam (minimum width) [m] [m]
Amritnagar Nega Jamehari 30 3.6 3.6 4.5 0.80 4.7
Amritnagar Nega Jamehari 30 3.6 3.6 6 0.60 5.4
Begonia Begonia 36 3.9 3.9 3 1.30 6
Amlai Burhar 30 4.5 4.5 5.4 0.83 4.5
Sendra Bansjora X 23 4.65 4.65 8.1 0.57 5.55
W. Chrimiri Main 90 5.4 5.4 3.75 1.44 6
Birsingpur Johilla top 129 7.5 7.5 3.6 2.08 6
Pure Kajora Lower Kajora 54 5.4 5.4 3.6 1.50 6
Pure Kajora Pure Kajora 56 4.95 4.95 3.6 1.38 6.45
Shankapur Jambad bottom 42 4.5 4.5 4.8 0.94 4.5
Collapsed

Ramnagar Begunia 70 2.85 2.85 1.8 1.58 3.15


Ramnagar Begunia 51 3 3 1.8 1.67 3.6
Kankanee XIII 160 19.8 19.8 6.6 3.00 4.2
Kankanee XIV 140 18.6 18.6 8.4 2.21 5.4
Jitpur XIV 450 18 34.5 3.6 6.50 6
Jitpur XIV 450 10.5 12 3.6 3.08 6
Jitpur XIV 450 12 21 3.6 4.13 6
Jitpur XIV 450 18 21 3.6 5.33 6
Jitpur XIV 450 18 25.5 3.6 5.83 6
Jitpur XIV 450 15 63 3.6 6.67 6
Jitpur XIV 450 18 30 3.6 6.25 6
Jitpur XIV 450 18 30 3.6 6.25 6
Jitpur XIV 450 30 43.5 3.6 5.67 6
Belampalli Ross 36 5.4 8.4 3 1.80 6
Nimcha Nega 48 9.9 9.9 6 1.65 6
Morganpit Salarjung 270 8.1 8.1 3 2.70 3.6
Ramnagar Ramnagar 75 9.9 9.9 2.7 3.67 6.6
Lacchipur Lower Kajora 38 7.2 7.2 5.1 1.41 3.9
N Salanpur X 30 9 9 5.1 1.76 6
Bankola Jambad top 102 10.1 10.1 4.8 2.10 2.4
Bankola Jambad top 75 6.3 6.3 3 2.10 4.2
Un-collapsed

Surakacchar G-I 106 16 16 3.5 4.57 4


Lacchipur Lower Kajora 38 18.3 18.3 5.1 3.59 4.2
Sripur Koithee 266 40 40 4.8 8.33 5
E Angarpathra XIII 30 6 6 2.1 2.86 6
Kargali incline Kathara 36 9.3 9.3 3.6 2.58 5.8
Jamadoba 6 & 7 XVI 80 5.8 5.8 2 2.90 5.5
Topsi Singharan 85 7 7 1.8 3.89 3.9
Jitpur XIV 450 21 39 3.6 7.60 6
Jitpur XIV 450 18 48 3.6 7.30 6
Jitpur XIV 450 19.5 30 3.6 6.50 6
Jitpur XIV 450 18 31.5 3.6 6.40 6
Jitpur XIV 450 18 42 3.6 7.00 6

212
11 | Coal Pillar Case Studies

11.2.5 Combined over the complete range of depths and of the un-
The aforementioned cases from Australia, South Africa, collapsed pillars below about 100 m depth.
the USA and India are shown combined in Figure The ranges covered by the combined coal pillar
11.5 for the collapsed and the un-collapsed pillars. A database are summarised in Table 11.11. As can be
general trend of increasing pillar width to height ratio seen in Figure 11.5, the combined data includes 15
with increasing depth can be seen, notwithstanding collapsed cases with w/h > 5 and three cases with w/h
the outlier cases previously discussed. The arbitrary approximately equal to 8. All the collapsed cases have
line shown is w/h = 2.0 + 0.0125 x depth. This does a w/h ≤ 8.
reasonable job in differentiating the collapsed pillars

FIGURE 11.5 The combined collapsed and un-collapsed database. The circled cases are the yielding pillars from Jitpur, India
and the highly rectangular pillars from Utah, USA. The bottom plots are subsets of the database focussing on the cases
above 200 m depth. The line shown is w/h = 2.0 + 0.0125 x depth.

213
ESTIMATING ROCK MASS PROPERTIES

TABLE 11.11 Summary of the combined coal pillar database


Case Number of cases Depth [m] w/h

Collapsed 162 19 – 488 0.6 – 8.0

Un-collapsed 507 13 – 594 1.0 – 16.7

11.2.6 Other one case, plotted within the bounds predicted by his
Gale (1998) demonstrated that the mechanical numerical modelling.
response of coal and surrounding strata defines pillar To further this research, the aforementioned
strength and argued that pillar strength formulae, such Australia, South Africa, USA and India databases are
as Equation 11.1 and Equation 11.2, over-simplify the superimposed onto this figure. Most the cases plot
problem. He plotted the results of numerical modelling with low vertical stress (Figure 11.7). However, in
as a graph of pillar strength and w/h to estimate the superimposing the cases it was necessary to assume
range of possible pillar strengths (Figure 11.6). In his that the vertical stress as calculated using the Tributary
work, Gale (1998) assumed a coal rock mass strength Area approximates the average pillar strength. While
of 6.5 MPa as indicated by the horizontal line labelled that assumption may be reasonable for collapsed
‘No Confinement Generated’ in Figure 11.6. pillars, it should under-estimate the strength of
Gale (1998) plotted onto this figure, field data from un-collapsed pillars.
34 cases from collieries in Australia, UK and the For that reason, the un-collapsed cases are removed
USA that had measured pillar stress values. All but and only the collapsed cases are shown in Figure 11.8.

FIGURE 11.6 Measured pillar stress from various pillar geometries as presented by Gale (1998). The shaded area shows the
range of possible pillar strengths

214
11 | Coal Pillar Case Studies

FIGURE 11.7 The Australian (blue), South African (green), USA (orange) and Indian (red) databases superimposed onto the
plot in Figure 11.6

The bulk of the 162 collapsed pillars plot within the


bounds suggested by Gale (1998). However, there are
42 South African, 13 USA and 5 Indian cases which lie
below 6.5MPa. Interrogating the combined database
the data indicates that, with the exception of the Jitpur
shaft pillars, these pillars are slender (w/h ≤ 4), which
according to Mark (2006) run the risk of ‘sudden,
massive collapse’. Specifically, the cases that plot below
the ‘No Confinement Generated’ line comprise:
• South Africa
„„ 29 pillars with w/h ≤ 3 in ‘normal’ coal
„„ 13 ‘weak’ coal pillars and with w/h ≤ 4
• USA
„„ 11 very slender pillars (w/h ≤ 2)
„„ 2 slender pillars (w/h ≤ 4)

FIGURE 11.8 Close-up of Figure 11.7 showing only • India


collapsed cases from Australia (blue), South Africa (green „„ 5 shaft pillars from Jitpur which is a weaker coal
triangles are ‘normal coal; green circles are ‘weak coal’), (Sheorey et al. 1987)
USA (orange) and Indian (red). The grey dashed line is
drawn at 3.15 MPa.
A value of 3.15 MPa for the ‘No Confinement
Generated’ line instead of the 6.5 MPa assumed by
Gale (1998) appears to be appropriate for the ‘weak’
coal pillars and the Jitpur cases (Figure 11.8). This is

215
ESTIMATING ROCK MASS PROPERTIES

in keeping with the average strength recommended by 1998; Galvin et al. 1999). The adopted value is mainly
van der Merwe (2003) for ‘weak’ coal. This still leaves dependent on the form of the empirical strength
the very slender pillars (w/h ≤ 2) outside the envelope. equation used, i.e. linear or power (see Equation 11.1
However, as described by York et al. (2000) thin pillars and Chapter 5.3).
are highly subject to geological structures – faults,
shears, joints. 11.3.2 South Africa
The fact that very slender pillars and pillars from Bieniawski (1968) collated the compressive strength
‘weak’ coal seams plot below the minimum strength results of 579 tests on 1 inch cubes of coal from 10
assessed by Gale (1998) suggests that the coal rock South African coal seams. The average results from
mass strength does play a role in pillar strength. the 10 sets of data ranged between 5000 to 8200 lb/
in2 (i.e. 34.5 to 56.5 MPa), with the overall average of
6054 lb/in2 (41.7 MPa).
11.3 UCS The cubes are not directly equivalent to UCS or
Coal is one of the few rocks that has been extensively triaxial core samples. Because of their shape and smaller
tested at various scales. The results of the testing size, the cubes are expected to give higher compressive
confirm that coal’s strength reduces as the sample size strengths than standard UCS samples. Townsend et
increases, e.g. Bieniawski (1968); Esterhuizen et al. al. (1977) tested over 200 cubes and cylinders with
(2010). But if the applicability of the typical methods areas of 3 to 16 in2 and found that small cylindrical
for estimating rock mass strength is to be assessed, e.g. samples (1.7” or 43 mm diameter) were typically 20 to
Hoek-Brown / GSI, then knowledge of the UCS of 30% weaker than cubical samples of the same cross-
coal is required. section. The difference between the cube and cylinder
As an aside, the UCS is simply a repeatable index of strength diminishes with increasing sample size. They
strength. In the case of coal, an organic rock giving it postulated that sample preparation caused the bulk
unique constituents and properties which tend to result of the difference. Masoumi (2013) also attributed the
in greater scatter in laboratory tests than other rocks decrease in strength he experimentally found with
(Hoek and Brown 1997), that repeatable index cannot core diameter smaller than about 50 to 65 mm, to the
be reliably obtained by uniaxial strength testing. relative proportion of surface damage during specimen
Instead, Medhurst and Brown (1998) recommend preparation. Hence, the expected compressive strength
estimating the UCS of coal from triaxial testing. of a standard UCS sample based on the 1-inch cube
tests could be approximately 20 to 30% weaker, i.e. in
11.3.1 Australia the range of 24 to 40 MPa with an average of 30 MPa.
Medhurst and Brown (1998) found a trend of decreasing Mathey and van der Merwe (2015) analysed 186
strength with increase in coal ‘brightness’. The ‘brighter’ BTS, 403 UCS, 198 triaxial tests from numerous seams
the coal, the higher its fixed carbon content and the from the South African coal fields. Their analysis did
more cleating it has. They estimated the UCS for coal not show significant differences between coals from
from the Moura mine in Queensland to be 32.7 MPa different seams. They found that:
for dull coal; 22.5 MPa for mid-bright coal; and 9.9
MPa for high-bright coal. They postulated that the
• a lognormal distribution with overall mean of 1.6
MPa and individual means for the coal seams of
laboratory samples of dull coal were essentially ‘intact’; between 1.05 – 1.89 MPa fits the BTS data
i.e. dull coal has negligible cleating; and that brighter
coals are ‘broken’ on cleating to various degrees. On • a normal distribution with overall mean of 23 MPa
and individual means for the coal seams of between
that basis, they derived Hoek-Brown parameters from 19 – 26 MPa fits the UCS data.
regression analysis of triaxial laboratory testing of 61 to
300 mm diameter core samples. The ratio between UCS and BTS implies mi = 23
For dull coal core samples, they proposed ‘intact / 1.6 = 14.4 with a range of 13.7 – 18.1. This is in
rock’ parameters of σc = 35 MPa, mi = 16.7, s = 1.0 and keeping with mi = 16.7 found by Medhurst and Brown
a = 0.5. For mid-bright coal mass, they proposed the (1998). As too is the mean UCS of 23 MPa compared
parameters of σc = 22.5 MPa, mb = 2.6, s = 0.052 and a with 22.5 MPa for mid-bright Moura coal.
= 0.65. The Moura coal is a medium rank high-volatile The range of estimated strengths for a 1 ft3 to
coal; a classification that places it in the middle of the 1 m3 cube of South African coal suggested by various
coalification series, i.e. anthracite – bituminous coal researchers is between 4.0 and 9.1 MPa (Taylor and
– sub-bituminous coal – lignite. Hence, the strength Fowell 2003). This range includes the ‘weaker’ coals
parameters of Moura coal might be reasonable ‘average’ from the Vaal Basin and Klip River, which van der
or ‘typical’ values. Merwe (1993) distinguishes as having a nominal
Increasing the sample scale to a nominal 1 m3 cube strength for 1 m3 cube of 4.5 MPa instead of the
of coal, the design compressive strength values adopted typical 7.2 MPa of Salamon and Munro (1967).
in Australia are in the order of 5.4 to 8.6 MPa, (Gale Salamon et al. (2006) quote slightly different values

216
11 | Coal Pillar Case Studies

of between 1.0 and 6.8 MPa, with overall averages reported UCS and triaxial testing of a further eight
of 4.6 to 5.4 MPa, again depending on the form of Indian coal samples that had σc ranging from 14.3 to
empirical equation used to estimate coal pillar strength. 43.7 MPa.
The results of in situ compressive testing of 0.3 m
11.3.3 USA cubes of coal were presented in a plot in Sheorey et
Holland (1964) summarised the compressive strength al. (1987) and showed a range of approximately 4 to
tests on 3-inch cube samples of coal as varying 13 MPa. Jawed et al. (2013) found that the empirical
between 1000 to 7000 psi (6.9 to 48.3 MPa) when power relationship proposed by Salamon and Munro
tested perpendicular to bedding and between 800 to (1967), best fits their data for slender pillars from
5000 psi (5.5 to 34.5 MPa) when tested parallel to Indian coal fields using a compressive strength of
bedding. Assuming that the equivalent cylindrical 7.6 MPa.
samples (i.e. 75 mm diameter) would be typically 20
to 30% weaker (Townsend et al. 1977), the implied 11.3.5 UK
compressive strength perpendicular to bedding is in In developing numerical models to study yielding
the range of 5 to 34 MPa. pillars Yavuz and Fowell (2001) derived the following
The individual tests of nearly two thousand UCS Hoek-Brown criterion parameters for the Parkgate coal
tests from 54 coal seams in the USA, range between seam: σc = 20 MPa, mi= 8.06, GSI = 46, mb = 1.17,
4 and 47 MPa (Mark and Barton 1997). Interestingly, s = 0.055 and a = 0.5. Yavuz and Fowell (2001) state
the overall average was found by to be 23.4 MPa, mi = 8.06 is the average for British coal.
which is a similar value to that of mid-bright coal from
Moura mine (22.5 MPa). 11.3.6 China
Published compressive strength test data of USA Coal samples comprising 50 mm diameter cylinders
coal was collated in Du et al. (2008) for mainly cube from the Dongsheng coal field, of the Yongli colliery
samples ranging between 1.5 to 54 inches. Their in Inner Mongolia, China were tested for strength in a
regression analysis of the different data sets implied the laboratory programme comprising five UCS, four BTS
strength of a 1-inch cube to be between 3238 to 9837 and 4 triaxial tests by He et al. (2016). The results were
psi (22.3 to 67.8 MPa) depending on the coal seam. (mean ± standard deviation): UCS = 21.7 MPa ± 5.5
Again assuming equivalent cylindrical samples (i.e. 25 MPa, BTS = 1.2 ± 0.2 MPa, and modulus, E = 2.9 GPa
mm diameter) (Townsend et al. 1977) this suggests a ± 1.2 GPa.
range of 15 to 47 MPa.
Esterhuizen et al. (2010) found the UCS of coal 11.3.7 Conclusions
varied between 16 and 40 MPa. They adopted the The UCS of coal is a difficult index to measure. It
following Hoek-Brown parameters for modelling coal is variable, broadly ranging from 5 to 45 MPa, which
pillars; σc = 20 MPa, mi = 9.8, mb =1.47, s = 0.07 and classifies coal as ‘weak’ to ‘medium strong’ according
a = 0.5. to ISRM (1978). However, the UCS ranges experienced
Mark and Barton (1997) concluded “ laboratory in Australia, South Africa, the USA, India, the UK
testing should not be used to determine coal strength” and China are remarkably similar as shown in Table
pointing out the strong influence that geological 11.12 with an overall ‘average’ of about 20 MPa. This
structure (i.e. cleating) has on coal strength. They is an interesting conclusion given the different geneses
recommended adopting a uniform value of 6.2 MPa of the coal deposits (see Section 11.3).
for the “coal strength” for US pillar designs (15). The design values adopted for the nominal 1 m3
Of interest, Unal (1983) gave RMR76 (= GSI) values of cube of coal in South Africa, India, Australia and the
between 40 to 70, typically 60, for 58 roof-fall cases USA are also relatively similar, typically in the order of
from underground coal mines in the USA. 4 to 7 MPa with an overall ‘average’ of about 6.5 MPa
(Table 11.13).
11.3.4 India Others have noted the coal strength paradox: “why
As part of their assessment of failed and stable coal a wide variation is commonly observed of the laboratory
pillars, Sheorey et al. (1987) determined the strength compressive strength of coal yet the mass strength is
of 25 mm cubes of coal and found it ranged from 19 to remarkably uniform between coal producing basins and
50 MPa. The lower values coming from the Jitpur XIV even between continents” (Poulsen and Adhikary 2013),
seam. Again assuming equivalent cylindrical samples Hoek-Brown criterion parameters that have been
(i.e. 25 mm diameter) (Townsend et al. 1977) this used in various locations as noted in the preceding
suggests a range of 13 to 35 MPa. Sheorey et al. (1989) sections are listed in Table 11.14.

(15)
Pillar design using the Analysis of Retreat Mining Pillar Stability (ARMPS) method

217
ESTIMATING ROCK MASS PROPERTIES

TABLE 11.12 Summary of coal UCS data


Equivalent core Country UCS [MPa] Reference
diameter [mm] range, mean

25 (a) South Africa 24 to 40, 30 Bieniawski (1968)

USA 15 to 47 Du et al. (2008)

India 13 to 35, 20 Sheorey et al. (1987)

50 USA 4 to 47, 23.4 (Mark and Barton 1997)

USA 16 to 40, 20 Esterhuizen et al. (2010)

Australia 10 to 33, 22.5 Medhurst and Brown (1998)

South Africa 19 to 26, 23 Mathey and van der Merwe (2015)

India 14.3 to 43.7, 29.7 Sheorey et al. (1989)

UK 20 Yavuz and Fowell (2001)

China 21.7 ± 5.5 He et al. (2016)

75 USA 5 to 34 Holland (1964)


(a)
assessed equivalent strengths of 25 mm diameter core from testing of 25 mm cubes

TABLE 11.13 Summary of compressive strength of 1m3 coal samples


Country Compressive strength [MPa] Reference

Australia 5.4 to 8.6 Gale (1998), (Galvin et al. 1999)

South Africa 7.2 (b) Salamon and Munro (1967)

4.0 to 9.1 Taylor and Fowell (2003)

4.6 to 5.4 Salamon et al. (2006)

USA 6.2 Mark and Barton (1997)

India 7.6 (b) Jawed et al. (2013)


(b)
strengths of 0.3 m cubes

TABLE 11.14 Summary of coal Hoek-Brown parameters that have been used by others
Country mi mb s a Reference

Australia 16.7 2.6 0.052 0.65 Medhurst and Brown (1998)

South Africa 13.7 – 18.1 - - - Mathey and van der Merwe (2015)

USA 9.8 1.47 0.07 0.5 Esterhuizen et al. (2010)

UK 8.06 1.17 0.055 0.5 Yavuz and Fowell (2001)

218
11 | Coal Pillar Case Studies

11.4 NUMERICAL MODELLING there are no issues with floor / roof failure and that
coal modulus can be assumed to be independent of
A series of numerical analyses have been carried out
confinement. This is a similar approach to that taken
using the three dimensional, finite element program
by other researchers, e.g. Duncan Fama et al. (1995).
Abaqus(16) and the two dimensional, finite element
For this reason, the rock mass modulus for the roof
program Phase2 (17).
and floor is assumed to be Em= 5000 MPa.
The first series models a generic layout of bord and
pillar workings in 3D and 2D (Section 11.5.2). The 11.4.1.2 In situ stress
intention is to confirm the ability of the numerical The in situ stresses are modelled as hydrostatic and are
modelling to match empiricism. Specific comparison considered to cover a broader range of stress conditions
with the results obtained using the empirical method within coal seams. This is based on several reasons.
of Galvin et al. (1999) is made. Mark and Gadde (2010) analysed a database of more
The next three series model the coal pillar cases than 350 stress measurements from underground coal
contained within the databases: mines and found that:
• The Australian database in 3D (Section 11.5.3) • horizontal stress within coal measure rocks, i.e.
• The South African database in 2D (Section 11.5.4) non-coal, comprises two components: one that is
• The USA database in 2D (Section 11.5.5). between 0.8 and 2.0 times the vertical stress and
one that is an excess stress of approximately 3.5 to
The overall objective of these analyses is to compare
10.5 MPa, which is independent of depth
the modelling results with proposed coal (rock mass)
strength criteria. • there is variability in the stress ratio across the
various coal fields throughout the world
11.4.1 Parameters • modulus was almost as powerful a predictor of the
stress as the depth.
11.4.1.1 Young’s modulus
Nemcik et al. (2006) also concluded that in layered,
The coal and rock mass are modelled as homogenous,
sedimentary coal measure strata there is a strong
isotropic, linearly elastic materials. This simplification
tendency for the stiffer rocks, i.e. non-coal, to carry
is apt for these studies as the intent is to assess the
more stress. High horizontal stresses where present
major and minor principal stresses acting in the
would therefore be carried by the non-coal rocks.
un-collapsed coal pillars and at the point of failure of
the collapsed coal pillars. Further, this study is focused on the stresses induced
within the coal pillars, which are predominantly
The equivalent mass modulus for the modelled coal
vertical. The horizontal stresses within the coal pillars
seam is assumed to be Em= 1000 MPa, based on the
are the confining stresses which at the edges of the
following studies on core samples and cube samples
pillars are zero. The same approach was followed
from 50 mm to 2 m.
by Duncan Fama et al. (1995) in their numerical
• In situ testing of coal pillars showed decreasing modelling of coal pillars.
modulus with increasing pillar size, varying from
approximately 5000 MPa (0.72x106 psi) for 50 mm
(2’) cubes to 2900 MPa (0.42x106 psi) for 127 mm
11.4.2 Generic cases
(5’) cubes (Bieniawski 1968) 11.4.2.1 3D
• an Heerden (1975) found an average modulus of
v
 A grid of coal pillars was modelled in Abaqus as shown in
4000 MPa for in situ testing of cubes up to 2 m in size Figure 11.9 with a 10 x 10 grid of 2 m high and variable
• A value of 2500 MPa was back-analysed by Duncan width square pillars to simulate bord and pillar workings.
Fama et al. (1995) in their numerical modelling The results of the modelling are presented as a series
• Esterhuizen et al. (2010) used numerical analysis to of contour plots showing stresses, pillar strength and
back-analyse a modulus for USA coal of 3000 MPa Factor of Safety (FoS). The predicted stresses within
• The range of modulus values obtained by Poulsen the pillars are taken as those acting along a horizontal
et al. (2014) from testing coal seam core and back- plane at mid-height through the coal pillars as shown
analysing the empirical approach of Galvin et al. in Figure 11.10.
(1999) ranged from 990 to 3000 MPa The ratio of Cp / Sp is adopted as the FoS where the
• Mathey and van der Merwe (2015) collate the laboratory coal pillar strength is estimated using Equation 11.4
testing of South African coal samples and quote an as proposed in Chapter 5.3 noting that coal under
average modulus of 4450 MPa for core samples. low confinement, such as in slender pillars or at the
The exact value of the modelled modulus is not edges of pillars in general, displays brittle behaviour
important for these studies if the relative stiffness of suggestive of a damage initiation spalling limit (DISL)
the coal and the surrounding rock mass is such that type approach.

(16)
Dassault Systems, version 2.0
(17)
RocScience, version 8.0

219
ESTIMATING ROCK MASS PROPERTIES

FIGURE 11.9 3D finite element model. In this example, a 2m thick coal seam is shown as the green coloured layer with 20
m x 20 m pillars (the overburden is not shown for clarity)

FIGURE 11.10 Contours of major principal stress, σ1 - top and FoS – bottom; on a horizontal plane through the mid-height of
the coal pillar. This example is for the case with 20 m square pillars at a depth of 500 m and UCS = 25 MPa. The bulk of the
coal pillar has a FoS ≈ 1.20, although a 1 m thick annulus has a FoS ≤ 0.5 and a further 2 m thick annulus has a FoS ≈ 1.0.

220
11 | Coal Pillar Case Studies

relationship proposed by Galvin et al. (1999) was used


to estimate Cp (Equation 11.1). i.e.
EQUATION 11.4

Two values of σc have been assumed; 10 and 25 MPa.


These approximately represent the bounds of “average” The Tributary Area Theory as defined by Equation
strengths of the UCS data presented in Section 11.4.4. 11.5 and Figure 11.11 is used to estimate Sp and hence
The FoS results are summarised in Table 11.15. FoS = Cp / Sp. The results are given in Table 11.16.
To compare the numerical modelling results, the
same 10 x 10 grid of 2 m high square coal pillars at 50,
100, 250 and 500 m depths was assessed following the
UNSW pillar design method. In this case, the linear EQUATION 11.5

TABLE 11.15 Summary of average FoS from Abaqus modelling of pillar geometries
PILLAR WIDTH 4 8 12 16 20
[m]
R=w/h 2.0 4.0 6.0 8.0 10.0
DEPTH [m] Average FoS for σc = 10 MPa & 25 MPa
50 0.90 & 1.50 1.90 & 3.30 2.40 & 4.20 3.00 & 5.10 3.30 & 5.70
250 0.45 & 0.60 0.80 & 1.20 1.10 & 1.40 1.20 & 1.60 1.20 & 1.65
500 0.40 & 0.45 0.70 & 0.80 0.85 & 1.05 0.95 & 1.15 0.95 & 1.20
1000 0.40 & 0.40 0.65 & 0.70 0.75 & 0.85 0.85 & 0.95 0.90 & 1.00

FIGURE 11.11 Definition of mining geometry following Galvin et al. (1999)

221
ESTIMATING ROCK MASS PROPERTIES

TABLE 11.16 FoS results from UNSW method


PILLAR WIDTH [m] 4 8 12 16 20
R = w/h 2.0 4.0 6.0 8.0 10.0
Pillar Strength
Cp [MPa] 7.37 11.88 16.38 20.89 25.40
Tributary Area [m2] 6.25 3.06 2.25 1.89 1.69
DEPTH [m] ESTIMATED PILLAR LOAD, Sp [MPa]
50 7.81 3.83 2.81 2.36 2.11
250 39.06 19.14 14.06 11.82 10.56
500 78.13 38.28 28.13 23.63 21.13
1000 156.25 76.56 56.25 47.27 42.25
DEPTH [m] FoS
50 0.94 3.10 5.83 8.85 12.04
250 0.19 0.62 1.17 1.77 2.40
500 0.09 0.31 0.58 0.88 1.20
1000 0.05 0.16 0.29 0.44 0.60

Equation 11.5 can be re-written to make pillar width The FoS predicted by the Abaqus modelling and by
a function of FoS. The equation becomes the following the UNSW method are plotted against each other in
cubic, noting that for a regular grid of square pillars, Figure 11.12. A good correlation does exist between the
w1 = w2, b1 = b2 and θ =90°. two approaches but it is not a one-to-one relationship.
Generally, the Abaqus model predicts:
• lower FoS than the UNSW method at 50 m depth
• about the same FoS at depths of 250 and 500 m
EQUATION 11.6 • a higher FoS at 1000 m depth.
The Abaqus modelling predictions seem to be more
The pillar widths and corresponding ratio R = w/h in keeping with the pillar dimensions using the UNSW
according to Equation 11.6 are tabulated in Table pillar design method for FoS = 1.62. This ability of
11.17 for: numerical analysis to match the empirical method is
• FoS = 1.0 (i.e. pillar strength = pillar load) explored a little further in the following section which
uses two-dimensional modelling.
• FoS = 1.62 and 2.09 which is equivalent to 1 pillar
failure in 100 and in 10,000 respectively according
to (Galvin 2006) when using the linear form of the
UNSW method, i.e. Equation 11.1.

TABLE 11.17 Recommended widths of 2 m high, square pillars following Galvin et al. (1999)
OVERBURDEN DEPTH SQUARE PILLAR WIDTH [m] & (R = w/h)
[m] FoS = 1.00 FoS = 1.62 FoS = 2.09
50 3.9 (1.95) 5.1 (2.57) 6.0 (2.99)
250 10.4 (5.21) 14.5 (7.27) 17.5 (8.75)
500 16.9 (8.47) 24.4 (12.22) 29.9 (14.97)
1000 28.9 (14.45) 43.2 (21.58) 53.8 (26.90)

222
11 | Coal Pillar Case Studies

10

6
Abaqus FoS

3
50 m
250 m
2 500 m
1000 m
1:1
1
1:1.62

0
0 1 2 3 4 5 6 7 8 9 10
UNSW FoS

FIGURE 11.12 Comparison of the FoS result from 3D finite element modelling assuming σc = 25 MPa, and the UNSW
equation for a 2 m high pillars at depths of 50, 250, 500 and 1000 m. The small dashed line offset from the FoS = 1.62 line is
drawn to better visually fit the data.

11.4.2.2 2D
Two dimensional, plane strain, elastic numerical The geometry is shown in Figure 11.13.
analyses were run using the finite element program • The bord widths are consistently 6 m.
Phase2 to model a regular grid of square, 2 m high, • The coal is modelled as an elastic material with
variable width, coal pillars at overburden depths of 50, modulus Em= 1000 MPa. The roof and floor were
100, 200, 250 and 500 m. modelled as five times stiffer to remove the issues
The pillar widths were chosen to reflect those with floor / roof failure.
calculated to give a FoS = 1.62 following the UNSW • In situ stress as hydrostatic.
method as per Table 11.18. For example, 5.2 m for 50
m; 15.4 m for 250 m; and 26.3 m for 500 m depth.

TABLE 11.18 Pillar widths recommended following Galvin et al. (1999) for FoS = 1.62
OVERBURDEN DEPTH [m] SQUARE PILLAR WIDTH [m]
UNSW
for FoS = 1.62
50 5.1
250 14.5
500 24.4
1000 43.2

223
ESTIMATING ROCK MASS PROPERTIES

FIGURE 11.13 Finite element model and material properties

11.4.2.3 Results
Although the two-dimensional nature of the analyses It is concluded that the numerical model predictions
models infinitely long pillars, the relatively wide pillars are broadly consistent with the UNSW method.
mean that vertical sections taken through their centres This result is in keeping with Maybee (2000) who
are broadly representative of the square pillars. The found that the stresses predicted from two-dimensional
stresses are taken as the values along a line at mid- finite element analyses agree with those from Tributary
height through the coal pillar as shown in Figure 11.14 Area Theory when the w/h ≥ 1.0.
and Figure 11.15. The results are summarised in Table
11.19 and Figure 11.16.

FIGURE 11.14 Results showing major principal stress for the case w =18.7 m and depth = 250 m

224
11 | Coal Pillar Case Studies

18

16

14

12

Stress [MPa]
10

4
Pillar Load
2
Cp
0
-10 -8 -6 -4 -2 0 2 4 6 8 10
Distance [m]

FIGURE 11.15 Pillar load and inferred strength (Cp) at mid height through the pillar for the case w = 18.7 m, overburden
depth of 250 m and σc = 10 MPa

TABLE 11.19 Comparison of FoS from Galvin et al. (1999) and the finite element modelling
OVERBURDEN SQUARE PILLAR FoS FoS Phase2
DEPTH [m] WIDTH [m] (w/h) UNSW σc = 10 MPa σc = 25 MPa
50 5.1 (2.6) 1.62 1.28 2.99
100 7.9 (3.9) 1.62 1.65 2.79
14.5 (7.3) 1.62 1.91 2.46
250
17.75 (8.75) 2.09 2.00 2.57
24.4 (12.2) 1.62 1.89 2.19
500
29.9 (15.0) 2.09 1.94 2.25

40

35

30

25
Cp (MPa)

20

15

10

5 sc 10MPa
sc 25MPa

0
0 5 10 15 20 25 30 35 40
Pillar Load (MPa)

FIGURE 11.16 Results of finite element modelling. The inferred pillar strength, Cp, is plotted against the predicted pillar load.
The modelling with UCS = 10 MPa is in keeping with the FoS predicted using the UNSW method with FoS = 1.62.

225
ESTIMATING ROCK MASS PROPERTIES

11.4.3 Australia • 
The damage threshold or systematic cracking
Based on the previous section which looked at generic (Equation 5.2), also with UCS*/UCS = 0.30 and σc
bord and pillar geometries and concluded that = 7.5 MPa, and mi = 16 (as per Table 11.14) and
numerical modelling is broadly consistent with the hence mb = 0.1296, s = 0.0081 and a = 0.25
UNSW pillar design method, this section models the
specific cases contained in the UNSW database.
Three dimensional Abaqus models were created of
the UNSW database of collapsed and un-collapsed EQUATION 11.8
pillars which are presented in Table 11.2. The predicted
major and minor principal stresses (σ1 and σ3) were • The fracture initiation criterion3 with σc = 6.5 MPa,
obtained at the edge, at the centre and as an average being the average value for 1m samples as given in
across the pillar along two horizontal planes within Section 11.4.6 is also shown
the pillar:
• 0.5 m below the top of the pillar
• at mid-height of the pillar.
An example of the results is presented in Figure 11.17. EQUATION 11.9
The results are also summarised in Figure 11.18 in
a plot of major versus minor principal stress. Because It appears that the damage threshold and fracture
the analyses are elastic, distinction is made between initiation criteria can differentiate between collapsed
the centre and the edge of the pillars, i.e. stress re- and un-collapsed coal pillars at low confining stresses.
distribution is not modelled. At higher confinement, the two criteria appear to
Strength envelopes representing the following under-predict pillar strength.
criteria are also shown: The Hoek-Brown parameters for mid-bright Moura
• Hoek-Brown criterion with parameters derived for coal do reasonably differentiate between collapsed and
mid-bright Moura coal, that is, σc = 22.5 MPa, mb = un-collapsed coal pillars.
2.6, s = 0.052 and a = 0.65 (see Section 11.4.1)  

• The fracture initiation criterion (Equation 11.4)


with UCS*/UCS = 0.30 and σc = 7.5 MPa which
was found by trial and error to give a better visual
fit to the data

EQUATION 11.7

226
11 | Coal Pillar Case Studies

FIGURE 11.17 Contours of pillar load, Sp or σ1 (top) and confining stress, σ3 (bottom). This example is for case FC13: 21 x 21
m, 4.9 m high pillars at a depth of 185 m with 6.1 m bord width

227
ESTIMATING ROCK MASS PROPERTIES

FIGURE 11.18 Predicted σ1 v σ3 for the Australian database plotted with fracture initiation envelopes for 6.5 and 7.5 MPa
compressive strength. The bottom plots show only the stresses from the middle height of the pillar. The left-hand plot does
not show the stresses for the centre of the pillars; the right does not show the stresses at the edge of the pillars. 

228
11 | Coal Pillar Case Studies

11.4.4 South Africa


A subset of approximately a quarter of the South
African database was created by selecting cases to
represent collapsed pillars from ‘normal’ and ‘weak’
coal and un-collapsed pillars at approximately 10 m
depth intervals. The ensuing subset of cases (Table
11.20) was modelled with the two dimensional, elastic
finite element program Phase2 .

TABLE 11.20 Subset of the South African coal pillar database used in modelling
Case Depth [m] Pillar Width Pillar Height w/h Bord Width Comment
[m] [m] [m]
n204 21 6.8 3.2 2.1 5.3 ‘Normal’
s41 30.5 4.6 3.7 1.2 7.6 ‘Normal’
n173 41 6.4 6.2 1 6.4 ‘Normal’
n188 51.5 6 3.9 1.5 6 ‘Normal’
n202 60 7 1.8 3.9 6 ‘Normal’
m149 90 7.5 4.8 1.6 6 ‘Normal’
n186 100 8.5 3.3 2.6 6.5 ‘Normal’
m168 165.5 15 5.9 5.2 5 ‘Normal’
s122 167.5 15.9 5.5 2.9 5.5 ‘Normal’
s67 184.5 15.9 5.5 2.9 5.5 ‘Normal’
s66 193 15.9 5.5 2.9 5.5 ‘Normal’
m170 205 17 4.9 3.5 6 ‘Normal’
s59 58 5.2 3.7 1.4 6.4 ‘Weak’
n182 70 12.5 2.9 4.3 5.5 ‘Weak’
n180 82 10 2.8 3.6 5 ‘Weak’
n194 96 12 6 2 6 ‘Weak’
n196 104 12 3 4 6 ‘Weak’
m159 108 10.6 3.2 3.3 6.5 ‘Weak’
m157 112 10.6 2.8 3.8 6.5 ‘Weak’
s60 152.5 12.2 4.9 2.5 6.1 ‘Weak’
2 21.5 6.1 2.7 2.2 6.7 Un-collapsed
23 30.5 6.4 3.6 1.7 6.4 Un-collapsed
83 41 6.4 2.0 3.2 6.4 Un-collapsed
146 46 7.6 4.9 1.6 6.1 Un-collapsed
142 50.5 6.1 1.4 4.5 6.1 Un-collapsed
32 61 7.6 4.0 1.9 6.1 Un-collapsed
89 76 7.6 4.6 1.7 6.1 Un-collapsed
38a 85 7.6 3.2 2.4 6.1 Un-collapsed
51 90 8.5 2.9 2.9 6.71 Un-collapsed
92 91 12.2 1.5 8 6.1 Un-collapsed
87 106.5 12.2 4.3 2.9 6.1 Un-collapsed
137 116 9.1 1.8 5 6.1 Un-collapsed
135 143.5 12.8 1.7 7.6 5.5 Un-collapsed
77a 161 14.9 3.0 5 6.3 Un-collapsed
77c 180.5 15.2 3.0 5.1 6.0 Un-collapsed
88 198 17.2 2.8 6.1 5.7 Un-collapsed
84 219.5 21.7 3.2 6.9 5.6 Un-collapsed

229
ESTIMATING ROCK MASS PROPERTIES

An example of the mesh geometry used is shown in • The damage threshold as per Equation 11.8 but
Figure 11.19. with σc = 4.5 MPa
• The coal is modelled as an elastic material with • The fracture initiation criterion with the ‘average’ of
modulus Em= 1000 MPa. The roof and floor were 6.5 MPa (Equation 11.9).
modelled as five times stiffer to remove the issues To make it easier to compare the results, they are
with floor / roof failure. also plotted separately for the collapsed cases in Figure
• In situ stress as hydrostatic. 11.22 and for the un-collapsed cases in Figure 11.23.
The stresses are taken as the values along a line Figure 11.22 shows most points that represent
taken at mid-height through the coal pillar as shown either average stresses or the stress acting at the edge
in Figure 11.20. The results are summarised in Figure of collapsed pillars lie above the proposed criteria.
11.21 in a plot of major versus minor principal stress Further, Figure 11.23 shows most points that represent
with the following strengths envelopes. either average stresses or the stress acting at the centres
• The fracture initiation criterion (Equation 11.7) of un-collapsed pillars lie below the proposed criteria.
with σc = 4.5 MPa which was found by trial and It therefore appears that the damage threshold and
error to give a better visual fit to the data fracture initiation criteria can differentiate between
collapsed and un-collapsed coal pillars reasonably well.

FIGURE 11.19 Example of the 2D FEM and material properties (w =15.9 m and depth = 193 m)

230
11 | Coal Pillar Case Studies

FIGURE 11.20 Results showing major principal stress for the case w =15.9 m and depth = 193 m

FIGURE 11.21 Predicted σ1 v σ3 for the South African database subset plotted with fracture initiation envelopes for 4.5 and
6.5 MPa compressive strength.

231
ESTIMATING ROCK MASS PROPERTIES

FIGURE 11.22 Predicted σ1 v σ3 of the South African collapsed pillar cases plotted with fracture initiation envelopes for 4.5
and 6.5 MPa compressive strength. The right-hand plot does not show the stresses for the centre of the pillars.

FIGURE 11.23 Predicted σ1 v σ3 of the South African un-collapsed pillar cases plotted with fracture initiation envelopes for
4.5 and 6.5 MPa compressive strength. The right-hand plot does not show the stresses at the edge of the pillars.

232
11 | Coal Pillar Case Studies

11.4.5 USA
A subset of approximately a quarter of the un-collapsed
cases (37 from 135 cases) and all the 35 collapsed cases
from the USA database was created. The ensuing
subset of cases (Table 11.21) was modelled with the
two dimensional, elastic finite element program Phase2 .

TABLE 11.21 Subset of the USA coal pillar database used in modelling (the last 12 collapsed cases are
those defined as massive collapses or rock burst and are shown in italics)
Case Depth Pillar Width [m] Pillar w/h Bord
[m] (minimum width) Height Width
State Mine [m] [m]
KY 20 442 15.2 x 9.1 9.1 1.8 5.0 6.1
KY 20 396 15.2 x 9.1 9.1 1.7 5.5 6.1
KY 42 320 15.5 x 11 11.0 1.8 6.0 5.8
KY 60 396 18.6 x 11 11.0 1.8 6.2 5.8
KY 74 326 22.6 x 16.5 16.5 3.7 4.5 5.5
KY 75 122 9.1 x 9.1 9.1 2.0 4.6 6.1
KY 91 244 15.2 x 12.2 12.2 2.3 5.3 6.1
KY 128 396 22.9 x 22.9 22.9 3.4 6.8 6.1
KY 128 488 29 x 24.4 24.4 3.2 7.6 6.1
KY 137 244 15.2 x 15.2 15.2 4.0 3.8 6.1
OH 92 158 9.8 x 7.7 7.7 1.8 4.2 5.5
OH 92 198 9.8 x 7.7 7.7 1.8 4.2 5.5
UT 59 305 18.3 x 12.2 12.2 2.0 6.2 6.1
VA 16 488 24.4 x 16.8 16.8 4.0 4.2 6.1
WV 2 91 3 x 12.2 3.0 2.4 1.3 6.1
Collapsed

WV 67 351 18.3 x 15.2 15.2 1.9 8.0 6.1


WV 79 130 12.2 x 3 3.0 2.4 1.3 6.1
WV 79 122 12.2 x 3 3.0 2.5 1.2 6.1
WV 81 76 12.2 x 3 3.0 2.1 1.5 6.1
WV 82 122 12.2 x 3 3.0 3.0 1.0 6.1
WV 133 122 12.2 x 3 3.0 1.5 2.0 6.1
WV 133 114 12.2 x 3 3.0 1.4 2.2 6.1
WV 134 76 13.7 x 3.8 3.8 2.1 1.8 6.1
CO 111 91 4.1 x 19.4 4.1 2.1 1.9 11.1
UT 11 168 12.2 x 12.2 12.2 5.3 2.3 6.1
WV 23 73 12.2 x 3 3.0 3.0 1.0 6.1
WV 23 73 18.3 x 3 3.0 3.0 1.0 6.1
WV 23 85 6.1 x 12.2 6.1 3.0 2.0 6.1
WV 23 75 3 x 12.2 3.0 3.0 1.0 6.1
WV 36 69 6.1 x 6.1 6.1 3.4 1.8 6.1
WV 57 53 12.2 x 3 3.0 3.0 1.0 6.1
WV 76 91 6.1 x 6.1 6.1 2.4 2.5 6.1
WV 82 84 12.2 x 3 3.0 2.9 1.1 6.1
WV 82 84 12.2 x 3 3.0 2.9 1.1 6.1
CO 116 472 18.3 x 15.2 15.2 2.6 5.9 6.1

233
ESTIMATING ROCK MASS PROPERTIES

TABLE 11.21 Subset of the USA coal pillar database used in modelling (the last 12 collapsed cases are
those defined as massive collapses or rock burst and are shown in italics) continued
Case Depth [m] Pillar Width [m] Pillar w/h Bord
(minimum width) Height Width
State Mine [m] [m]
PA 99 64.0 5.2 x 5.2 5.2 1.3 4.0 5.5
PA 113 76.2 12.5 x 12.5 12.5 1.5 8.2 5.8
KY 25 88.4 14.9 x 15.2 14.9 1.9 7.8 6.1
WV 118 114.3 18.3 x 12.2 12.2 1.6 7.4 6.1
VA 112 121.9 19.2 x 12.2 12.2 1.4 8.4 6.1
UT 5 137.2 12.5 x 12.2 12.2 1.4 8.9 6.1
KY 114 152.4 18.3 x 18.3 18.3 1.4 13.3 6.1
OH 92 161.2 9.8 x 7.7 7.7 1.8 4.2 5.5
WV 33 170.7 18.3 x 18.3 18.3 3.6 5.1 6.1
OH 92 181.4 9.8 x 7.7 7.7 1.8 4.2 5.5
KY 61 189.0 12.5 x 15.2 12.5 2.2 5.6 6.1
OH 92 204.4 11.3 x 9 9.0 1.8 4.9 5.5
WV 6 210.3 21.6 x 21.3 21.3 1.6 13.2 6.1
WV 83 219.5 14.9 x 15.2 14.9 1.6 9.2 6.1
VA 112 233.2 19.2 x 12.2 12.2 1.4 8.4 6.1
Un-collapsed

VA 64 243.8 24.4 x 18.3 18.3 1.3 14.1 6.1


UT 4 259.1 20.4 x 18.3 18.3 2.1 8.6 6.1
KY 100 274.3 9.1 x 9.1 9.1 1.6 5.8 6.1
VA 64 280.4 18.3 x 18.3 18.3 1.3 14.1 6.1
KY 100 289.6 9.1 x 6.1 6.1 1.3 4.6 6.1
KY 100 304.8 9.1 x 9.1 9.1 1.4 6.3 6.1
VA 64 313.9 18.3 x 18.3 18.3 1.3 14.1 6.1
KY 100 320.0 12.2 x 12.2 12.2 1.4 8.4 6.1
WV 40 333.5 18.3 x 27.4 18.3 2.0 9.0 6.1
KY 21 341.1 12.2 x 15.2 12.2 1.5 8.3 6.1
KY 21 353.6 18.3 x 21.3 18.3 1.6 11.1 6.1
KY 131 365.8 12.5 x 12.5 12.5 1.7 7.5 5.8
VA 88 396.2 24.4 x 21.3 21.3 1.5 14.0 6.1
VA 88 411.5 17.4 x 15.2 15.2 1.5 10.0 6.1
KY 32 423.7 18.3 x 15.2 15.2 1.6 9.6 6.1
KY 61 432.5 18.3 x 24.4 18.3 2.2 8.2 6.1
KY 61 489.5 18.3 x 24.4 18.3 2.2 8.2 6.1
UT 35 502.9 34.4 x 9.1 9.1 3.0 3.0 6.1
VA 8 518.2 18.3 x 18.3 18.3 1.7 10.9 6.1
KY 77 536.8 18.3 x 18.3 18.3 1.7 10.9 6.1
UT 3 585.2 11.3 x 30.5 11.3 2.9 3.9 6.1
UT 35 594.4 34.4 x 9.1 9.1 3.0 3.0 6.1

234
11 | Coal Pillar Case Studies

The same mesh geometry as in the UNSW and • Hoek-Brown criterion with parameters derived for
South African cases is used. USA coal, that is, mb = 1.47, s = 0.07 and a = 0.5 (see
The stresses are taken as the values along a line Section 11.4.1) and σc = 6.0 MPa
taken at mid-height through the coal pillar as shown
in Figure 11.24.
The results are summarised in Figure 11.25 and in
Figure 11.26 as plots of major versus minor principal Again, it appears that the damage threshold and
stress with following strength envelopes. fracture initiation criteria can differentiate between
• The fracture initiation criterion (Equation 11.7) collapsed and un-collapsed coal pillars reasonably
with σc = 6.0 MPa which was found by trial and well; see the right-hand plots of Figure 11.25 and
error to give a better visual fit to the data Figure 11.26. That the criteria capture the coal pillars
• e damage threshold as per Equation 11.8 but
Th
 that were recorded as either massive collapse or rock
with σc = 6.0 MPa burst lends support to the view that coal can be
• The fracture initiation criterion with the ‘average’ of thought of as a brittle rock under low confinement
6.5 MPa (Equation 11.9) (see Chapter 5.3.1).

FIGURE 11.24 Results showing major principal stress for the case w =9.1 m and depth = 396 m

235
ESTIMATING ROCK MASS PROPERTIES

FIGURE 11.25 Predicted σ1 v σ3 of the USA collapsed pillar cases plotted with fracture initiation envelopes with 6 and 6.5
MPa compressive strength. he right- hand plot does not show the stresses for the centre of the pillars. The pillars defined as
‘massive collapse and rock burst’ are in the lighter shade.

FIGURE 11.26 Predicted σ1 v σ3 of the USA un-collapsed pillar cases plotted with fracture initiation envelopes with 6 and
6.5 MPa compressive strength. The right-hand plot does not show the stresses for the centre of the pillars.

236
11 | Coal Pillar Case Studies

11.4.6 India
In the same fashion as the other databases, the Indian Again, it appears that the damage threshold and
database (Table 11.10) was modelled with the two fracture initiation criteria can differentiate between
dimensional, elastic finite element program Phase2. The collapsed and un-collapsed coal pillars reasonably well
stresses along a line taken at mid-height through the coal (right hand plots of Figure 11.27).
pillar are summarised in Figure 11.27 in a plot of major
versus minor principal stress with the strength envelopes:
• The fracture initiation criterion (Equation 11.7)
with σc = 5.0 MPa which was found by trial and
error to give a better visual fit to the data
• The damage threshold as per Equation 11.8 but
with σc = 5.0 MPa
• The fracture initiation criterion with the ‘average’ of
6.5 MPa (Equation 11.9)

FIGURE 11.27 Predicted σ1 v σ3 plotted of the Indian collapsed (top) and un-collapsed (bottom) pillar cases with failure
envelopes. The right-hand plots do not show the stresses for the centre of the collapsed pillars and the edge of the un-
collapsed pillars. The yielding pillars from Jitpur, India are circled.

237
ESTIMATING ROCK MASS PROPERTIES

11.5 DISCUSSION • The fracture initiation criterion


The preceding analyses show the predictions from finite
element modelling are consistent with the empirical
relationships for sizing coal pillars. The results of the • 
The Hoek-Brown criterion with parameters
analyses also show that the damage threshold criterion recommended by:
(Equation 5.2) and a fracture initiation criterion (such „„ Medhurst and Brown (1998) for mid-bright
as Equation 11.4) can differentiate between collapsed Moura coal, i.e. mb = 2.6, s = 0.0052 and
and un-collapsed coal pillars, particularly at low a = 0.65. Recall that Moura coal classifies in
confining stresses. At higher confinement, the criteria the middle of the coalification series and hence
may under-predict the coal pillar strength. these parameters might be reasonable ‘average’ or
One of the issues in using these criteria is determining ‘typical’ values
the UCS of coal, which as discussed in Section 11.4, is „„ E sterhuizen et al. (2010) for USA coal, i.e. mb =
not a straightforward value. Though an average value 1.47, s = 0.07 and a = 0.5 which are similar to
of 6.5 MPa for the compressive strength of 1m3 cubes those for UK coal used by Yavuz and Fowell
was used, the following slightly different values were (2001), i.e. mb = 1.17, s = 0.055 and a = 0.5.
found by trial and error to give visually better fits to • A maximum confined strength, e.g. Barton
the data. (1976); Bieniawski (1968); van Heerden (1975),
• 7.5 MPa for the 18 collapsed and 15 un-collapsed which is taken to correspond to 1 as per
Australian cases Singh et al. (2011). The maximum confined
• 5.0 MPa for the 86 collapsed and 337 un-collapsed strength is 0.6 based on the normalised
South African cases deviatoric plot shown in Figure 11.32.
• 6.0 MPa for the 35 collapsed and 135 un-collapsed It is remarkable that so many cases are differentiated
USA cases by the criteria. This is particularly so considering the
• 5.0 MPa for the 23 collapsed and 20 un-collapsed rather generalised approach adopted in assigning UCS
Indian cases. and estimating the various parameters for the criteria.
The ability of the criteria to differentiate is perhaps
The above values were used to normalise the
better shown in Figure 11.30 and Figure 11.31, which
principal stresses predicted by the numerical modelling
show collapsed and un-collapsed coal pillars separately.
of the coal pillar databases. The normalised principal
Further, as the criteria capture the USA coal pillars that
stresses along a line through the mid-height of the coal
were recorded as either massive collapse or rock burst
pillars are plotted in Figure 11.28.
lends support to the view that coal can be thought of
Figure 11.29 plots the same data but only the average as a brittle rock under low confinement.
stresses. Also shown in Figure 11.29 are:
• The damage threshold criterion

FIGURE 11.28 Predicted principal stresses normalised with the various assumed compressive strengths. The left graph shows
collapsed pillars (closed symbols) and the right un-collapsed (open symbols) pillars. The symbols are also used to represent
average stress along the mid-height plane of the pillar (square symbols); stresses at the pillar edge (triangles) and stresses in
the centre of the pillar (circles). The light blue diamonds are “massive collapse or rock burst” USA coal pillars. The dashed line
represents hydrostatic stress.

238
11 | Coal Pillar Case Studies

FIGURE 11.29 Predicted principal stresses normalised with the various assumed compressive strengths. The criteria –
damage threshold, crack initiation, Hoek-Brown, and a constant deviatoric stress – are shown. The bottom graph is a close-up
of the stresses under low confinement. Closed symbols are collapsed pillars, open symbols are un-collapsed pillars. The light
blue diamonds are “massive collapse or rock burst” USA coal pillars.

239
ESTIMATING ROCK MASS PROPERTIES

FIGURE 11.30 Predicted principal stresses normalised with the various assumed compressive strengths for the collapsed
pillars. The criteria – damage threshold, crack initiation, Hoek-Brown, and a constant deviatoric stress – are shown. The
bottom graph is a close-up of the stresses under low confinement. The light blue diamonds are “massive collapse or rock
burst” USA coal pillars.

240
11 | Coal Pillar Case Studies

FIGURE 11.31 Predicted principal stresses normalised with the various assumed compressive strengths for the un-collapsed
pillars. The criteria – damage threshold, crack initiation, Hoek-Brown, and a constant deviatoric stress – are shown. The
bottom graph is a close-up of the stresses under low confinement.

241
ESTIMATING ROCK MASS PROPERTIES

FIGURE 11.32 Predicted principal stresses normalised with the assumed UCS values. This is the same data as shown in
Figure 11.29 but as deviatoric stress. The light blue diamonds are “massive collapse or rock burst” USA coal pillars.

242
11 | Coal Pillar Case Studies

11.6 CONCLUSIONS The combined database of 162 collapsed pillars.


Tributary load against w/h ratio. The line corresponding
The different approach used for estimating coal pillar
to the pillar strength as calculated using the UNSW
strength to the typical method for other rock masses
linear equation is also shown with a FoS = 1.85.
has been examined.
A large subset of the combined database – 96
The coal pillar databases from Australia, South
collapsed and 89 un-collapsed – was analysed with
Africa, the USA and India were collated to generate
2D and 3D finite element analyses to assess the likely
a combined database comprising 162 collapsed and
principal stresses within the coal pillars. These stresses
507 un-collapsed pillars covering depths from 13 to
were initially plotted on principal stress plots along
594 m with w/h ratios between 0.6 and 16.7. One of
with a fracture initiation criterion of the form:
criticisms of the UNSW coal pillar design method is
that its database only contains one collapse case with
w/h > 5, e.g. Colwell (2010). The combined database
now contains 15 collapsed cases with w/h > 5. This is
shown in Figure 11.33.

Collapsed Pillars
25

20

FoS = 1.85

15
Pillar Load [MPa]
Tributary Area

Australia
10 South Africa
South Africa - weak
USA - squeeze
USA - massive
5
USA - local bursting
India
Cp UNSW linear

0
0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0 18.0
w/h = R

FIGURE 11.33 The combined database of 162 collapsed pillars. Tributary load against w/h ratio. The line corresponding to
the pillar strength as calculated using the UNSW linear equation is also shown with a FoS = 1.85.

243
ESTIMATING ROCK MASS PROPERTIES

The results showed that this form of criterion could


differentiate between collapsed and un-collapsed coal
pillars at least at low confining stresses. At higher
confinement, the criterion appears to under-predict
the coal pillar strength. The difficulty in using this
criterion was assigning an appropriate UCS value to
the coal.
The principal stresses predicted in the finite element
modelling were then normalised with the following
slightly different values which were found by trial and
error to give visually better fits to the data.
• 7.5 MPa for the 18 collapsed and 15 un-collapsed
Australian cases
• 5.0 MPa for the 86 collapsed and 337 un-collapsed
South African cases
• 6.0 MPa for the 35 collapsed and 135 un-collapsed
USA cases
• 5.0 MPa for the 23 collapsed and 20 un-collapsed
Indian cases.
The combined database was represented on the same
plot together with the following strength criteria:
• The damage threshold criterion

• The fracture initiation criterion

• The Hoek-Brown criterion with

&
mb s a

• A maximum confined strength

It is remarkable that these criteria could differentiate


so many cases, especially considering the rather
generalised approach taken here in assigning UCS
and estimating the various parameters for the criteria.
That the criteria capture the USA coal pillars that were
recorded as either massive collapse or rock burst lends
support to the view that coal can be thought of as a
brittle rock under low confinement (see Chapter 5.3.1).

244
12 | Rock Mass Strength

12 ROCK MASS
STRENGTH

12.1 OUTLINE 12.1.1 Variability


As discussed in Chapter 5.6, the Hoek-Brown rock Such is the nature of geology that there are several steps
mass criterion was initially based on very limited test in the empirical process followed in this work that can
work from Panguna andesite, the largest samples of lead to uncertainty in the derived relationships. Minor
which had 22.5” (572 mm) diameters. sources of potential error include testing conditions
(Pells 1993) and the choice of regression method used
In this chapter, an extensive database of large
to curve-fit the data (Mostyn and Douglas 2000). The
scale hard rock failures is assembled from literature.
major sources of potential error though stem from the
The database comprises mine pillars in igneous,
values assigned to GSI and to the Disturbance Factor, D.
metamorphic and sedimentary rocks and both small
and large diameter core testing. In each case the rock 12.1.1.1 GSI
mass quality, or data that could use to assess rock mass GSI can be calculated from a trimmed version of RMR
quality, had been recorded. (Hoek 1994; Hoek and Brown 1997), though Hoek
This chapter collates published accounts of: (2007) found the correlation to be problematic for
• in situ stress failures of hard rock from underground poor quality rock masses and recommended that the
mining in metamorphic rocks, limestone, iron ore GSI be estimated directly by means of its charts. The
and sandstone GSI can also be estimated from a quantified chart as
suggested by several authors, e.g. (Cai et al. 2004; Hoek
• laboratory UCS and TXL testing where rock mass et al. 2013; Russo 2009; Sonmez and Ulusay 1999).
quality has been interpreted.
The importance of estimating GSI by more than one
A database of UCS and TXL from testing carried method is re-stated (Bertuzzi et al. 2016; Cai et al.
out in the 1990s of rock core from the Ok Tedi mine 2004; Hoek et al. 2013; Russo 2009). It is particularly
in Papua New Guinea is also studied to assess the important that a geological model that includes the
variation in measured strengths with rock quality as problem scale is developed in order to estimate GSI
gauged by GSI. (Bertuzzi et al. 2016; Carter and Marinos 2014). The
The objective of this collation of rock mass strength GSI values of the data used in this work, which cover
is to supplement the stress failure data from the a broad range of scales from laboratory to mine pillar,
Hawkesbury Sandstone tunnels in Chapter 10 and reflect the scale of the individual data sets, i.e. the size
from the coal pillar database in Chapter 11. This effect has been accounted for by appropriate selection
chapter then uses this combined data to empirically of GSI. Hence, it is possible to combine these data.
derive relationships between GSI and the parameters Several researchers consider there is a limit to the
mb, s and a of the Mostyn-Douglas modified Hoek- range of GSI values over which the Hoek-Brown
Brown criterion. The criterion is curve-fitted to criterion may be applied (Brown 2008; Carter et al.
the data to ensure that the derived relationships are 2007; Carvalho et al. 2007; Diederichs 2007; Goricki
independent of the standard Hoek-Brown/GSI (Hoek et al. 2006). Brown (2008) opines ‘special care must
et al. 2002). be exercised at low values of GSI of below about 30’ or
It is assumed that a = 0.5 for intact rock to keep ‘ σc < 15 MPa’. The approach adopted is to plot the
consistent with the Hoek-Brown equation. It is noted data as assessed but cautiously interpret the results as
though that better fits for intact rock can be found recommended by Brown (2008).
by modifying a, e.g. (Carter et al. 2007; Carter et
al. 2008; Carvalho et al. 2007; Mostyn and Douglas 12.1.1.2 D
2000; You 2011) however this becomes problematic In relation to D, Hoek et al. (2002) noted that “a large
when extending into rock mass behaviour. number of factors can influence the degree of disturbance

245
ESTIMATING ROCK MASS PROPERTIES

in the rock mass surrounding an excavation and that it Lunder (1994) reported the rock masses had typical
may never be possible to quantify these factors precisely”. RMR values of between 60 and 85; which classifies as
Hoek and Diederichs (2006) elaborated further by Good to Very Good rock. This is consistent with the
stating that D “will vary for each application, depending typical GSI values quoted by Martin and Maybee
upon the excavation and loading sequence for the (2000) for Canadian Shield rocks of 40, 60 & 80.
particular structure being designed.” The approach is to A summary of the database is presented in Table
follow the guidelines provided in Hoek et al. (2002) to 12.1. Lunder (1994) presented the pillar database in
assign values of D to the individual datasets. two ways.
Firstly, he plotted the normalised principal stress σ1/
σc as a function of the pillar width to pillar height ratio
12.2 MINE PILLARS w/h as shown in Figure 12.1. This is the traditional
12.2.1 Metamorphic way of plotting pillar load performance data. Lunder
(1994) estimated σ1 either with numerical modelling
A database comprising 178 hard rock pillars from
or Tributary Area Theory.
underground mining of strong to extremely strong,
mainly massive sulphide deposits of Canada was Secondly, he plotted σ1/σc versus a term he defined
compiled by Lunder (1994). The UCS for the as ‘the average pillar confinement’ (Cpav) which is
massive sulphide rock ranges from 70 to 315 MPa; approximately equal to the average principal stress
for the quartzite between 85 to 168 MPa and for the ratio (σ3/σ1) across the mid-height centreline of the
metasediments is 240 MPa. The database excluded pillar. Lunder’s plot is shown in Figure 12.2. Also,
pillars affected by major geological structures and recall that the same plane was used, i.e. the mid-height
covered a wide range of pillar sizes: centreline of the pillar, to assess the stresses in the coal
pillars in Chapter 11.
• Pillar width: 1.9 to 45 m
• Pillar height: 2.4 to 53 m
• Average pillar stress: 25 to 128 MPa.

TABLE 12.1 Hard rock pillar database collated by Lunder (1994)


Source Number of pillars Rock mass
Failed Marginal Stable
Massive sulphide
Lunder, 1994 18 11 2
Canada
Massive sulphide
Hudyma, 1988 12 9 26
Canada
Quartzites
Hedley & Grant, 1972 3 2 23 Elliot Lake, Ontario
Canada
Subtotal
33 22 51
(Canada cases)
Massive sulphide
Metasediments
Von Kimmelman et al, 1984 29 11 7
Selbi-Phikwe mine
South Africa
Metasediments
Sjoberg, 1992 5 4
Sweden
Massive sulphide
Krauland & Soder, 1987 13 Black Angel mine
Greenland
Massive sulphide
Brady, 1977 1 2 Mt Isa, Qld,
Australia
TOTAL 68 50 60

246
12 | Rock Mass Strength

1.2
Lunder
Hudyma
Von Kimmelman
1.0 Hedley & Grant
Sjoberg
Krauland & Soder
0.8 Brady
σ1 / σc

0.6

0.4

0.2

0.0
0.0 1.0 2.0 3.0 4.0 5.0
w / h

FIGURE 12.1 Ratio of σ1 v σc versus w/h for the hard rock pillar database collated by Lunder (1994). Stable pillars are shown
as open symbols; marginal pillars by lightly shaded symbols; failed pillars by solid symbols.

1.2
Lunder
Hudyma
Von Kimmelman
1.0
Hedley & Grant
Sjoberg
Krauland & Soder
0.8 Brady
σ1 / σc

0.6

0.4

0.2

0.0
0.0 0.1 0.2 0.3 0.4 0.5
Average pillar confinement (approx. σ3/σ1)

FIGURE 12.2 Ratio of σ1 v σc versus average pillar confinement for the hard rock pillar database collated by Lunder (1994).
Stable pillars are shown as open symbols; marginal pillars by lightly shaded symbols; failed pillars by solid symbols.

247
ESTIMATING ROCK MASS PROPERTIES

Lunder calculated Cpav with an empirical formula That is, a plot of σ1/σc against σ3/σc . The ratio σ3/σc
he derived from numerical modelling to be valid for can be calculated by multiplying the ‘average pillar
w/h < 4.5 (Equation 12.1)(18). Figure 12.3 plots this confinement’, Cpav, by σ1/σc, viz:
relationship.

EQUATION 12.1
The resultant principal stress plot is shown in Figure
It would be very useful to represent Lunder’s 12.4 for massive sulphide and Figure 12.5 for quartzite
database on a non-dimensional principal stress plot. and metasediment.

0.5

0.45

0.4

0.35

0.3
Cpav

0.25

0.2

0.15

0.1

0.05

0
0 1 2 3 4 5 6 7 8
w/h

FIGURE 12.3 Relationship between Cpav and w/h following Equation 12.1 from Lunder (1994). The relationship is only valid
for w/h < 4.5.

FIGURE 12.4 Non-dimensional principal stress σ1 v σc versus σ3 v σc for the hard rock pillars in the database collated by
Lunder (1994). The plots are of the 141 massive sulphide pillars: the 60 failed pillars are shown as solid symbols (left); 44
marginal pillars as lightly shaded symbols (middle); and 37 stable pillars as open symbols (right). A visually fitted Hoek-Brown
criterion is also shown.

(18)
 under (1994) also proffered an equation to estimate the Factor of Safety (FoS) for hard rock pillar design based on this concept of ‘the average pillar
L
confinement’.

248
12 | Rock Mass Strength

FIGURE 12.5
Non-dimensional principal stress σ1 v σc versus σ3 v σc for the hard rock pillars in the database collated by Lunder (1994).
The top plots are of the 28 quartzite pillars: the 3 failed pillars are shown as solid symbols (left); 2 marginal pillars as lightly
shaded symbols (middle); and 23 stable pillars as open symbols (right). The bottom plots are of the 9 metasediments pillars:
the 5 failed pillars are shown as solid symbols (left); 4 marginal pillars as lightly shaded symbols (middle); there are no stable
pillars in the database (right). A visually fitted Hoek-Brown criterion is also shown.

249
ESTIMATING ROCK MASS PROPERTIES

The failed hard rock pillar data collated by Lunder the published ranges of mi for these rock types
(1994) spans ranges of: (Hoek 2001).
• 70 ≤ σc ≤ 315 MPa
• 55 ≤ GSI ≤ 75 12.2.2 Limestone
• 0 ≤ σ3 /σc ≤ 0.2 and 0.18 ≤ σ1 /σc ≤ 1.04. The USA’s National Institute for Occupational
The Hoek-Brown criterion shown in Figure 12.4 Safety and Health (NIOSH) collated pillar geometry,
and Figure 12.5 was visually fitted to the data with stability conditions and RMR’89(19) data at 92 sites from
parameters: 34 underground mines. The mines use room and pillar
methods to excavate the Palaeozoic-aged (Ordovician
• mb = 3, s = 0.04, a = 0.5 for the massive sulphide to Carboniferous) limestone deposits of the Interior
pillars
Plains and the Appalachian Highlands of the USA
• b = 7, s = 0.07, a = 0.5 for the quartzite pillars
m (Esterhuizen et al. 2008).
• mb = 1.5, s = 0.015, a = 0.5 for the metasediments The limestone deposits in the Interior Plains are
pillars.
generally sub-horizontal and those in the Appalachian
The resulting curves do not differentiate all the data Highlands are folded and faulted. Only mines
but they do a reasonable job. where the strata dip 10° or less are included in the
The Hoek-Brown parameter mi for intact massive NIOSH database. Table 12.2 summarises the mining
sulphide ore is estimated as 172 / 12 ≈ 14 based on the geometries. The limestone is a very strong to extremely
values UCS = 172 MPa ± 4 MPa and tensile strength strong rock as indicated by the UCS testing summarised
= 12 ± 12 MPa given in Lunder (1994). Recall that mi in Table 12.3. The RMR’89 values are in a relatively tight
can be estimated as σc/|σt| (see Chapter 2). range of between 65 and 85, with the overall mean and
Lunder (1994) does not provide tensile strength standard deviation as shown in Table 12.4 and Figure
data for the metasediment or quartzite. Hence, the mi 12.6. This classifies the limestone as Good to Very
for the metasediments (σc = 240 MPa) and quartzite Good Rock which is in keeping with the fact that only
(σc = 85 to 168 MPa) was estimated at 12 – 22 (17) half of the mines install regular roof reinforcement in
and 17 – 23 (20), respectively, to be consistent with 10 to 15 m wide openings (Esterhuizen et al. 2011).

TABLE 12.2 Mining dimensions of NIOSH limestone mine database (Esterhuizen et al. 2008)
Depth [m] Pillar width Pillar height w/h
(w) [m] (h) [m]

Average 117 13.1 11.1 1.41

Minimum - 23 – 671 4.6 - 21 4.8 – 38 0.29 – 3.52


Maximum

TABLE 12.3 UCS testing of NIOSH limestone mine database (Esterhuizen et al. 2008)
Number of Average Range
Representative Formations
tests [MPa] [MPa]

Burlington, Salem, Galena-


Very Strong 50 88 44 – 143
Platteville

Camp Nelson, Monteagle,


Very to Extremely Strong Plattin, Vanport, Upper Newman, 100 135 82 – 207
Chickamauga

Extremely Strong Loyalhanna, Tyrone 32 219 151 – 301

(19)
1989 version of RMR

250
12 | Rock Mass Strength

Several thousand stable pillars and 18 cases of


 ummary of RMR of NIOSH limestone
Table 12.4 S
mine database (Esterhuizen et al. 2011) failed individual pillars in otherwise stable layouts are
included in the NIOSH database. The UCS of limestone
Parameter Ratings forming these 18 failed pillars varied between 150
Mean (standard
deviation) and 215 MPa, i.e. extremely strong rock. Many of the
instabilities were related to pillars that were initially
Strength of intact rock stable during development but became unstable during
UCS [MPa] or after benching, i.e. as w/h decreased (Esterhuizen et
11.6 (1.79)
al. 2006).
RQD and spacing of defects
[Defect frequency]
Excluding the six cases where large defects
23.0 (3.91) intersected the pillars as reported by Esterhuizen et al.
(2008), the 12 pillars failed at (see Figure 12.7):
Defect length 4.0 (1.23)
• stresses between 0.11 and 0.17 times the UCS, with
Condition of

Separation 5.5 (1.14) one exception at 0.08 UCS


defects

Roughness 4.7 (1.27) • w/h between 0.4 to 0.9 with one exception at w/h
= 1.7. The pillar for this one exception failed by
Infill (gouge) 5.7 (0.82)
progressively spalling.
Weathering 6.0 (0.10)

Groundwater 14.4 (2.1)

RMR’89 75 (range 65 – 80)

GSI = RMR’89- 5 70 (range 60 – 75)

FIGURE 12.6 Distribution of RMR in the NIOSH database; the minimum and maximum values are shaded (Esterhuizen et al.
2011)

251
ESTIMATING ROCK MASS PROPERTIES

FIGURE 12.7 Summary of pillar layouts and the 18 single failed pillars from Esterhuizen et al. (2008)

The data in Figure 12.7 that represent the 12 failed Essentially the limestone pillars failed in unconfined
pillars and several of the stable pillars adjacent to the compression at stresses approximately 11% to 17% of
line labelled the ‘boundary of current experience’ the UCS. That is when σ3 = 0, σ1 ≈ 0.11σc to 0.17σc.
was digitised. The ‘average pillar confinement’ was Substituting this into the Hoek-Brown criterion
then calculated for these pillars using the empirical
equation of Lunder (1994) in Equation 12.1. In
this way, principal stress pairs (σ3/σc , σ1/σc ) were
found. The resultant principal stress plot is shown in it simplifies to:
Figure 12.8.

252
12 | Rock Mass Strength

0.6

0.5

0.4
σ1 / σc

0.3

0.2

pillar failed by
progressive spalling

0.1
Failed

stable
0.0
-0.1 0.0 0.1 0.2 0.3
σ3 / σc

FIGURE 12.8 Ratio of σ1 v σc versus σ3 v σc for the failed limestone pillars (solid symbols) and a selection of the stable
limestone pillars (open symbols) collated by Esterhuizen et al. (2008).

253
ESTIMATING ROCK MASS PROPERTIES

12.2.3 Iron ore (Grgic et al. 2001). It is interpreted that the chemical
Hosni et al. (2015) assessed the conditions of nine degradation reduces the strength of the iron ore.
pillar collapses of the underground iron ore mines in The following is calculated.
the Lorraine area of France between 1997 and 2009. • The σ1/σc ratio shown in Table 12.5 based on a UCS
The oolitic-type iron ore occurs in 3 to 7 m thick value of 19 MPa, which Grgic et al. (2003) quotes
layers separated by limestone and sandstone – ‘marls’ as being the representative short-term UCS of the
(argillite, siltite and carbonates) (Grgic et al. 2003). iron ore core samples. Note that it is likely that
The mines were closed in 1997. this is the maximum UCS value at the time of pillar
Hosni et al. (2015) estimated the total stress acting failure.
on the pillars at the times of their collapse using the • The Cpav for these pillars using Equation 12.1
Tributary Area Theory, to range between 8 and 12 • The σ3/σc ratio by multiplying Cpav by σ1/σc.
MPa as listed in Table 12.5. Chemical degradation The resultant principal stress pairs are plotted in
of the iron ore with time led to the pillar collapses Figure 12.9.

TABLE 12.5 Summary of iron ore pillar collapses (Hosni et al. 2015)
Event Pillar dimension [m] Bord Pillar
Depth
w/h width Stress, σ1 σ1 / σc
Height Width Length [m]
[m] [MPa]
Mou_1997 3.5 12 12 3.4 6 120 8 0.42
Aub_96_02 5 12 14 2.4 6 170 10.9 0.57
Affl_01 7 24 24 3.4 6 255 12 0.63
Doma_01 5 27 27 5.4 6 245 11 0.58
Cru_1977 3.8 11 25 2.9 / 6.6 6 148 10.9 0.57
Rue Dante 5 12 49 2.4 / 9.8 8 170 9.9 0.52
Roc_2008 3 10 37 3.3 / 12.3 3 190 8 0.42
Ron_1999 2.5 6 85 2.4 / 34.0 6 140 9 0.47
Ang_2009 5 11 40 2.2 / 8.0 6 175 9.4 0.49

0.8

0.7

0.6

0.5
σ1/σc

0.4

0.3

0.2

0.1 Failed
mb = 0.7, s = 0.01, a = 0.55
0.0
-0.1 0.0 0.1 0.2 0.3
σ3/σc

FIGURE 12.9 Ratio of σ1 v σc versus σ3 v σc for the failed iron ore pillar database collated by Hosni et al. (2015). The Hoek-
Brown envelopes were curve-fitted.

254
12 | Rock Mass Strength

The Hoek-Brown envelopes were curve-fitted. The inferred failure envelope mb = 0.7, s = 0.01, a =
The failed iron ore pillar data spans ranges of: 0.55 (Figure 12.9) implies an unconfined compressive
strength of the iron ore pillars of approximately 10% of
• σc ≈ 19 MPa the UCS. That is when σ3 = 0, σ1 ≈ 0.10σc and
• GSI ≈ 40 – 60 as estimated from published photos
for a blocky to very blocky rock mass with fair to sa=0.10
good surface conditions of fractures (Figure 12.10)
• 0.1 ≤ σ3/σc ≤ 0.2 and 0.42 ≤ σ1/σc ≤ 0.63.

FIGURE 12.10 Photos of Lorraine iron ore mines. Top from Geoderis (2008); bottom from website accessed on 15/06/15:
https://c1.staticflickr.com/5/4121/4777055460_bc878f0850_z.jpg

255
ESTIMATING ROCK MASS PROPERTIES

12.2.4 Sandstone 0.40


Krauland et al. (1989) described a full-scale pillar
test, conducted between 1983 and 1988, in the Stable
Laisvall lead-zinc-silver mine, northern Sweden. The
orebody, hosted in flat-lying, Cambrian to Devonian Failed
aged, quartzitic sandstone interlayered with clayey
sandstones, was mined using the room and pillar 0.30
mining method until 2001.
The sandstone is interbedded with thin shale
partings and overlayed by Cambrian-aged schist. The
sandstone’s UCS varies between 130 and 290 MPa,
with a typical value of 210 ± 40 MPa (Soder and

σ1 / σc
0.20
Krauland 1989).
The full-scale test was conducted on nine pillars in
one area of the mine. The objective was to determine
the stress level in the pillar at failure to serve as input
to pillar design (Idris et al. 2015). The pillars were
subjected to increasing stresses by decreasing the 0.10
cross-sectional area of the pillars through blasting six
approximately 0.4 m thick slices, thereby reducing the
widths and lengths of the pillars in each of the mining
steps (Soder and Krauland 1989).
The pillars were initially 4.6 m high, 7.4 m wide
and 8.1 m long which gives a minimum w/h ratio 0.00
of 1.6. Pillar failure was defined to have occurred 0.0 0.1 0.2
when fractures were visually assessed to extend to the σ31 /σc
central parts of the pillar. Failure generally occurred
after six slices were blasted, i.e. when the pillars
were approximately 4.5 – 6.1 m wide and 5.3 – 6.9m FIGURE 12.11 Ratio of σ1 v σc versus σ3 v σc for the the
long. These dimensions give a minimum w/h of Laisvall mine test pillars after Krauland et al. (1989)
1.0 – 1.2.
Soder and Krauland (1989) report the stress in the
pillars before the test had been 18.6 ± 3.0 MPa and
at failure the stress had increased to 24.2 ± 1.4 MPa.
12.3 LABORATORY SAMPLES
Based on these measurements, the calculated rock
12.3.1 Stripa granite
mass UCS as reported by Soder and Krauland (1989)
and in Edelbro et al. (2007) is 19.8 ± 1.4 MPa. A large core sample of granite (quartz monzonite)
1 m diameter by 2 m long was recovered by slot drilling
The following is then calculated: from the Stripa mine in Sweden for laboratory testing
• σ1/σc ratio based on a UCS value of 210 MPa (Thorpe et al. 1980). The uniaxial compressive strength
• Cpav for these pillars using Equation 12.1 test result of this large core sample gave a UCS of 7.55
• σ3/σc ratio by multiplying Cpav by σ1/σc. MPa and a corresponding modulus of 52.3 GPa.
The resultant principal stress pairs are plotted in The granite core was mapped as shown in Figure
Figure 12.11. 12.12. Two main joint sets were identified along with
It is clear from this figure that the pillars failed other random, short fractures. The RMR’89 of the
essentially in unconfined compression at stresses granite core as listed in Table 12.6 is estimated from
approximately 11% of the UCS. That is, when σ3 = 0, the information in the appendix of Thorpe et al. (1980).
σ1 ≈ 0.11σc and hence: A GSI of 53 – 65 is estimated using the equation GSI =
RMR89 – 5 (Hoek 1994; Hoek and Brown 1997) which
sa=0.11
is in keeping with the visual interpretation of the
mapping Figure 12.12; that is a very blocky rock mass
Idris et al. (2015) calculated a mean GSI of 59 and with good surface conditions of fractures. It is noted
range of 51 to 64, for the sandstone pillars by adopting that estimates of GSI should also be made independent
the relationship between block volume, Vb , and joint of RMR, e.g. (Bertuzzi et al. 2016; Hoek 2007; Hoek
condition, Jc , suggested by Cai and Kaiser (2007). et al. 2013).

256
12 | Rock Mass Strength

TABLE 12.6 RMR for the large granite core estimated from the data in Thorpe et al. (1980)
Parameter Maximum Description Rating
Rating
Intact rock strength 15 80 to 200 MPa 8 – 14
RQD 20 1.45 m Length of core and 22 fractures 11
Cumulative length of core ≥ 100mm = 0.83 m
Estimated RQD = 57
Joint spacing 20 Two main continuous joint sets with random 6–9
discontinuous fractures
Average spacing = 70 mm to 300 mm
Joint roughness 6 Smooth to slightly rough 1–3
Joint infill thickness 6 Less than 1 mm thick 4
Joint fill strength 6 Calcite and chlorite coatings 4
Joint weathering 6 Joint walls are generally unweathered, hard rock 5–6
Joint length 6 Continuity less than 1 – 3 m 4
Groundwater 15 Dry 15
RMR’89 100 58 – 70
GSI RMR’89 - 5 53 – 65

FIGURE 12.12 Fracture traces mapped in the large diameter Stripa granite core (Thorpe et al. 1980). The top figure shows
all the fractures mapped. The bottom figure shows the principal fractures and the locations of instrumentation. This bottom
figure suggests a GSI = 50 – 60.

257
ESTIMATING ROCK MASS PROPERTIES

Table 12.7 and Table 12.8 reproduce the results The unconfined compressive strength for the 1 m
of laboratory testing that was carried out on 52 mm diameter Stripa granite core, which is estimated to have
diameter core of the Stripa granite (Thorpe et al. a GSI = 50 - 65, is 7.55 MPa. This is approximately 7.5
1980). This data is presented in the principal stress plot / 175 = 4% of the intact UCS. That is, when σ3 = 0,
in Figure 12.13. The average UCS from this data is σ1≈ 0.04σc and hence we can write:
approximately 175 MPa. s a=0.04
The Hoek-Brown criterion parameters calculated
using the program RocLab(20) for the intact Stripa
granite data shown in Figure 12.13 are: σc = 175
MPa, mi = 25, s = 1.0 and a= 0.5. Intact granite has
GSI = 100.

TABLE 12.7 Laboratory testing of 52 mm core of 600


the Stripa granite from Thorpe et al.
(1980)
Specimen σ3 [MPa] σ1 [MPa] E [GPa]
A natural, healed fracture
was observed in sample

S1 6.90 247.8 58.81

S2 3.45 210.5 56.48


500
S3 0 82.2 53.16

S4 0 156.6 74.61

S5 6.90 151.2 46.06


400
T1 2.6 0 -

T2 4.1 0 -

S6 9.7 0 -
σ1 [MPa]

S7 7.1 0 - 300
S8 6.90 329.4 48.04
Intact

S9 3.45 264.2 52.86

S10 0 208.2 52.81

S11 0 178.3 54.54


200

Intact
TABLE 12.8 Summary of other laboratory testing previous
of conventional 52 mm core of Stripa 100
granite from Thorpe et al. (1980) naturally fractured
σ3 [MPa] σ1 [MPa] E [GPa]
1m diameter core
13.3 ± 1.4 0 -
HB: sc=175 mi=25
15.0 ± 1.8 0 -
0
0 214 ± 24 52.3 ± 6.5 -20 0 20 40 60 80 100120140160
0 208 ± 31 69.4 ± 6.6
σ3 [MPa]
5 309 ± 10 75.4 ± 1.8

10 372 ± 26 77.2 ± 0.9 FIGURE 12.13 Principal stress plot for the Stripa granite
core after Thorpe et al. (1980). The Hoek-Brown envelopes
20 470 ± 6 82.2 ± 2.2 were curve-fitted.

30 530 ± 14 83.2 ± 0.6

(20)
RocScience, version 1.033

258
12 | Rock Mass Strength

12.3.2 Lac du Bonnet granite That is, for the Lac du Bonnet granite rock mass,
Martin (1993) reported results of various testing carried the unconfined compressive strength, or as Martin
out on samples of Lac du Bonnet Granite, including (1993) called the crack damage stress, is approximately
direct tensile tests, as part of the AECL’s URL. Of 0.8σc . Recall that crack initiation typically starts at
interest in relation to the work in this chapter is the approximately 0.3σc to 0.5σc (Chapter 5.2).
strength testing of core samples ranging from 33 to Martin et al. (1999) quotes RMR ≈ 100 for the Lac
300 mm diameter. Martin (1993) concluded that the du Bonnet granite rock mass. Diederichs (2007) gives
Hoek-Brown criterion parameters for the intact Lac du similar values:
Bonnet granite core samples are: σc = 210 MPa, mi = • mi = 30, s = 1 and a = 0.5 for intact
28.9, s =1.0, a = 0.5 (Figure 12.14). • GSI = 85, mb= 17.6, s = 0.19 and a = 0.5 for rock
Martin (1993) also concluded that: mass (Table 6.6)
• Beyond approximately 140 mm diameter, the
unconfined compressive strength is relatively
constant at about 0.8 of the UCS
• The effect of scale on the compressive strength of
Lac du Bonnet granite is not very significant.

FIGURE 12.14 Principal stress plot (left) and UCS versus core diameter (right) for the Lac du Bonnet granite from Martin
(1993)

259
ESTIMATING ROCK MASS PROPERTIES

12.3.3 Artificially jointed granite For the artificially jointed granite samples, Alejano
Alejano et al. (2015) carried out triaxial testing of et al. (2015) assessed GSI = 82 and 65 as being
intact and artificially jointed granite cylinders of 54 appropriate for the three and five joint samples,
mm diameter. They assessed the GSI for these cylinders respectively (Figure 12.15). These GSI values together
as 100, 82 and 65 for the intact, three and five joint with σc = 118 MPa and mi = 41.4, correspond to the
samples, respectively (Figure 12.15). following Hoek-Brown parameters as calculated using
the program RocLab (20).
For intact granite, they assessed GSI = 100,
σc = 118 MPa and mi = 41.4. This value of mi may • For GSI = 82, mb = 21.8, s = 0.1353 and a = 0.500
be questionable though, as they did not include tensile • For GSI = 65, mb = 11.9, s = 0.0205 and a = 0.502.
testing and limited confining stress to between 0.5 and However, the Mostyn-Douglas method, which
12 MPa, which corresponds to a maximum σ3 /σc ratio allows the parameter a to vary, to curve-fit the Hoek-
of 0.1. The curve-fit mi value may therefore not be the Brown envelope to the data is also used (Mostyn and
best estimate based on the recommendations of: Douglas 2000). The resulting envelopes provide a
• Hoek and Brown (1980), who recommended a better fit to the data (Figure 12.16):
confining stress range between σt and 0.5σc • For GSI = 65, mb = 9.53, s = 0.222 and a = 0.712
• the recent review of Read and Richards (2015) who • For GSI = 82, mb = 25.63, s = 0.105 and a = 0.5.
recommend a confining stress range between σt and
0.3σc.

FIGURE 12.15 Samples of artificially jointed granite tested by Alejano et al. (2015) who assessed the GSI values of 100, 82
and 65 (left to right).

260
12 | Rock Mass Strength

FIGURE 12.16 Results of laboratory testing and inferred strength envelopes for the samples of artificially jointed granite
tested by Alejano et al. (2015). The Hoek-Brown envelopes were curve-fitted.

261
ESTIMATING ROCK MASS PROPERTIES

12.3.4 Quartzite
Douglas (2002) reported the results of four sets of The GSI predicted mb, s and a fits the data very well
triaxial testing on drill-core sized samples of quartzitic except for the data corresponding to GSI = 50. The
sandstone carried out by Habimana et al. (2002). Mostyn & Douglas modified criterion fits all the data
The four sets of quartzite classified as GSI = 15, very well with the following parameters.
25, 50 and 80 and with intact Hoek-Brown criterion • For GSI = 15, mb = 2.07, s = 0.0242 and a = 0.999
parameters, σc = 90 MPa, mi = 9. The Habimana et • For GSI = 25, mb = 1.56, s = 0.0011 and a = 0.653
al. (2002) data is presented as principal stress plots in • For GSI = 50, mb = 4.83, s = 0.006 and a = 0.617
Figure 12.17. • For GSI = 80, mb = 4.82, s = 0.0277 and a = 0.5.
The the Hoek-Brown strength criterion was curve- The triaxial testing of quartzite data spans ranges of:
fitted to the data using RocLab (20) to predict mb, s and
a from GSI. A curve-fit for the Hoek-Brown strength
• σc = 90 MPa
criterion as modified by Mostyn & Douglas (Mostyn • 15 ≤ GSI ≤ 80
and Douglas 2000) was also made. • 0 ≤ σ3 /σc ≤ 0.25 and 0.006 ≤ σ1 /σc ≤ 1.1.

FIGURE 12.17 Principal stress plots for the quartzite after Habimana et al. (2002) as given in Douglas (2002). The Hoek-
Brown envelopes were curve-fitted.

262
12 | Rock Mass Strength

12.3.5 Phyllite
Habimana et al. (2002) also presented strength curves • For GSI = 5, mb = 0.8, s = 0.0031 and a = 0.854
for drill-core sized samples of phyllite, which has been • For GSI = 15, mb = 0.804, s = 0.0021 and a = 0.689
digitised to obtain a few data points, see Figure 12.18. • For GSI = 25, mb = 1.16, s = 0.0043 and a = 0.780
These points are then presented in Figure 12.19 with
the GSI derived Hoek-Brown criterion parameters
• For GSI = 35, mb = 0.996, s = 0.0011 and a = 0.626.
The triaxial testing of phyllite data spans ranges of:
using the values for intact phyllite given in Habimana
et al. (2002) of σc = 70 MPa, mi = 8. • σc = 70 MPa
Again, the the Hoek-Brown strength criterion as • 5 ≤ GSI ≤ 35
modified by Mostyn & Douglas (Mostyn and Douglas • 0 ≤ σ3/σc ≤ 0.21 and 0.01 ≤ σ1/σc ≤ 0.57.
2000) was curve-fitted. The Mostyn & Douglas
modified criterion perfectly fits the data with the
following parameters.

FIGURE 12.18 Digitised points of the phyllite failure envelopes presented in Habimana et al. (2002)

263
ESTIMATING ROCK MASS PROPERTIES

FIGURE 12.19 Principal stress plots for the phyllite, after Habimana et al. (2002). The Hoek-Brown envelopes were curve-
fitted.

264
12 | Rock Mass Strength

12.3.6 Other This therefore implies that the back-analysed


Martin et al. (1999) back-analysed published accounts unconfined compressive strength (i.e. when σ3 = 0)
of stress-induced failures surrounding drives in various of the various rock masses assessed by Martin et al.
rock types and showed that the depths of failure could (1999) is:
be reasonably matched by adopting Hoek-Brown
criterion parameters for the rock mass of mb = 0 and
s = 0.11.
Recalling that the Hoek-Brown criterion is:
The cases are summarised in Table 12.9.

When m = 0 and s = 0.11, this equation simplifies to:

TABLE 12.9 Summary of the UCS for the intact rock and for the rock mass of the cases back-analysed by
Martin et al. (1999).
Back-analysed
UCS
Source, Rock type RMR’89 GSI UCS of rock mass
[MPa]
[MPa]

AECL’s URL, Canada Granite 224 ≈100 80 – 95 75


(85)

Donkin-Morien coal Sedimentary 36 85 80 12


mine, Canada

Metasediments 100 66 61 33

Canadian mine Granite gneiss 240 70 65 80

265
ESTIMATING ROCK MASS PROPERTIES

12.4 OK TEDI CORE TESTING 12.4.1 GSI


A rock testing programme was undertaken in the early The description, photographs and sketches were used
1990s by Ok Tedi Mining Limited on core from 10 to classify the hornfels and limestone cores with the
boreholes from its mine in Papua New Guinea. The GSI system chart. The data was then grouped into
testing, which was carried out by the Snowy Mountains GSI values of 50-59, 60-69, 70-79, 89-90 and 90-100.
Engineering Corporation (SMEC), included detailed This process is in keeping with that used by Hoek and
descriptions, sketches and photographs of samples Brown (1980) to link their strength criteria with the
prepared for UCS and TXL testing. Figure 12.20 RMR for the Panguna andesite samples.
shows an example of the testing reports. The results of triaxial testing for the monzonite
The rock types tested comprised: and monzodiorite were too few and too erratic to be
reliably used.
• Hornfels (metamorphosed sandstones and siltstones)
– 66 TXL and 44 UCS tests
• Limestone (including limestone breccia) – 39 TXL
and 19 UCS
• Monzonite – 9 TXL and 26 UCS (13 of which were
dry)
• Monzodiorite – 8 TXL and 6 UCS.
The diameter of the tested core was generally 61 mm
though some tests were conducted on 50 and 83 mm
diameter core. Confining pressures up to 15 MPa were
used for the TXL testing.

FIGURE 12.20 Example of the core description, photographs and sketches provided by SMEC with the test results.

266
12 | Rock Mass Strength

12.4.2 Intact rock properties The average UCS values quoted above and the UCS
Samples which classified with a GSI ≥ 90 are considered and TXL test data were used to calculate the Hoek-
to represent intact rock for the purposes of estimating Brown envelope parameters mi, s and a following the
the UCS. The results are presented as frequency curves methodology of Mostyn and Douglas (2000). The
in Figure 12.21 with the following averages. curve-fitted envelopes are shown in Figure 12.22 and
with the parameters summarised in Table 12.10.
• Hornfels, average UCS = 137 MPa
• Limestone, average UCS = 75 MPa.

TABLE 12.10 Hoek-Brown parameters for the limestone and hornfels test results
Rock Number of
tests mb s a r2
GSI

50 - 59 3 4.00 0.2711 0.820 0.6829

60 – 69 11 6.28 0.2714 0.522 0.5862


Limestone

70 – 79 14 9.56 0.6415 0.626 0.9090

80 – 89 0 - - - -

90 – 100 24 14.98 1.0000 0.500 0.8510

50 - 59 8 5.06 0.0160 0.500 0.7148

60 – 69 4 7.53 0.1230 0.831 0


Hornfels

70 – 79 35 8.51 0.2181 0.654 0.7360

80 – 89 45 14.4 0.3687 0.500 0.2380

90 – 100 21 39.1 1.0000 0.500 0.4110

100% 100%

90% 90%
Ok Tedi Hornfels Ok Tedi Limestone
80% 80%

70% 70%

60% 60%

50% 50%
All All
40% 40%
GSI>90 GSI>90
30% 30%
Average Average
(GSI>90) (GSI>90)
20% 20%

10% 10%

0% 0%
0 50 100 137 150 200 0 20 40 60 75 80 100 120
UCS [MPa] UCS [MPa]

FIGURE 12.21 Summary of UCS test results for hornfels and limestone core. The averaged values of those cores assessed to
have a GSI ≥ 90 to be the UCS of the rock type.

267
ESTIMATING ROCK MASS PROPERTIES

FIGURE 12.22 Summary of UCS and TXL test results for hornfels and limestone core with Hoek-Brown envelopes fitted to
the data following the methodology of Mostyn and Douglas (2000). The test data was plotted for GSI ranges of 50-59, 60-
69, 70-79, 80-89 and 90-100. There was no data for limestone in the GSI = 80-89 class and only four UCS data for hornfels
GSI = 60-69 to plot a reliable graph.

268
12 | Rock Mass Strength

12.5 DISTURBANCE FACTOR TABLE 12.11 Broad subjective categories of the


As discussed in Chapter 5.6, the introduction of a Disturbance Factor, D, for the rock
Disturbance Factor, D, enabled Hoek et al. (2002) mass databases
to combine the then relationships between the Hoek- D Database
Brown criterion parameters m and s and the RMR of
0 Sandstone pillar test – controlled blasting
‘undisturbed’ and ‘disturbed’ rock mass (Brown and Habimana sandstone and phyllite – drill core
Hoek 1988). “D is a factor which depends upon the Lac du Bonnet granite from URL – controlled
degree of disturbance to which the rock mass has been blasting
subjected by blast damage and stress relaxation. It varies Sedimentary – TBM
from 0 for undisturbed in situ rock masses to 1 for very Ok Tedi core samples
disturbed rock masses” (Hoek et al. 2002). 0– Hawkesbury Sandstone – mainly mechanical
Hoek et al. (2002) report that experience in the design 0.8 excavation but some blasting
of open pit slopes and the back-analysis of monitoring Coal pillars – mainly mechanical excavation
or controlled blasting
data from an underground cavern in Taiwan, led them Stripa granite – blasted tunnel but slot
to the guidelines for estimating the factor D which are drilled core
reproduced in Figure 12.23. They noted that “ it is clear Artificially jointed granite – man-made
that a large number of factors can influence the degree of fractures
disturbance in the rock mass surrounding an excavation 0.8 Hard rock pillars – production blasting
and that it may never be possible to quantify these factors Limestone pillars – production blasting
precisely.” Hoek and Diederichs (2006) elaborated Mines - metasediments, gneiss – production
further by stating that D “will vary for each application, blasting
depending upon the excavation and loading sequence for Iron ore pillars – production blasting
the particular structure being designed.”
Each of the preceding rock mass data sets was
subjectively catergorised for the appropriate D
following the guidelines in Figure 12.23. The resulting
broad categories are shown in Table 12.11 with the
reasons for assigning the D values. It is stressed that
these categories are, by definition, imprecise.

FIGURE 12.23 Guidelines for estimating disturbance factor, D, for tunnels from Hoek et al. (2002) and repeated in Hoek
(2012)

269
ESTIMATING ROCK MASS PROPERTIES

12.6 GSI RELATIONSHIPS derive relationships between GSI and the parameters
mb, s and a of the Mostyn-Douglas modified Hoek-
The data sets of rock mass in situ stress failures presented Brown criterion.
in this chapter are summarised in Table 12.12 together The relationships between mb and GSI, s and GSI and
with those from the Hawkesbury Sandstone tunnels a and GSI, are shown in Figure 12.24 to Figure 12.26,
in Chapter 10 and from the coal pillars in Chapter 11. respectively. The relationships between mb , a, GSI and D
This combined database is then used to empirically are shown in Figure 12.27. As stated in Section 12.2.1,

TABLE 12.12 Summary of the data used


No. of Intact rock Rock mass
data
Rock Mass sets of Hoek- σc GSI Maximum stress Hoek-Brown parameters
‘failures’ Brown [MPa]
mi σ3/σc σ1/σc mb mb/mi s a

Hawkesbury 9 12 30 – 35 65 – 0.1 0.45 4.11 0.343 0.0357 0.501


Sandstone 80
Chapter 10 (70)

Coal 162 8.1 – 18 4–4 40 – 2.5 3.9 1.47 0.08 0.07 0.5
Chapter 11 (16.7) 20 70 -0.18
(60) (0.088)

Massive 60 failed 14 70 – 55 – 0.2 0.68 3 0.214 0.04 0.5


sulphide 44 315 75
Lunder (1994) marginal (65)

Quartzite 3 failed 17 – 23 85 – 55 – 0.2 1.04 7 0.30 0.07 0.5


Lunder (1994) 2 marginal (20) 168 75 –0.41
(65)

Metasediments 5 failed 12 – 22 240 55 – 0.04 0.41 1.5 0.07 0.015 0.5


Lunder (1994) 4 marginal (17) 75 –0.13
(65) 0.088

USA limestone 12 - 44 – 60 – ≈0 0.17 - sa = 0.11 to 0.17


Esterhuizen et 300 75
al. (2008) (67)

Laisvall 9 - 130 – 51 – ≈0 0.11 - sa = 0.11


sandstone 290 64
Krauland et al. 210 (59)
(1989)

Lorraine iron 9 - 19 55 – 0.2 0.63 0.7 - 0.01 0.55


ore 70
Hosni et al. (62)
(2015)

Stripa Granite 1 25 175 53 – 0 0.04 - sa = 0.04


Thorpe et al. 1 m3 core 68 (55)
(1980)

Lac du Bonnet 1 28.9 210 80 – ≈0 0.8 17.6 0.609 0.19 0.5


Granite 95
Martin (1993) (85)

Sedimentary 1 - 36 80 ≈0 0.11 0 0 0.11 0.5


Martin (1993)

Metasediments 1 - 100 61 ≈0 0.11 0 0 0.11 0.5


Martin (1993)

Gneiss 1 - 240 65 ≈0 0.11 0 0 0.11 0.5


Martin (1993)

Granite core 19 41.4 118 82 0.1 2.38 28.53 0.689 0.105 0.5
Alejano et al.
(2015) 12 41.4 118 65 0.1 2.38 9.53 0.230 0.222 0.712

270
12 | Rock Mass Strength

there is a limit to the range of GSI values over which


the Hoek-Brown criterion may be applied. The figures
plot the data as assessed with the area GSI < 30 shaded
to indicate that in this area the relationships should be
used cautiously as recommended (Brown 2008).
The equations derived from this data are presented in
Table 12.13.

TABLE 12.12 Summary of the data used continued


No. of Intact rock Rock mass
data
Rock Mass sets of Hoek- σc GSI Maximum stress Hoek-Brown parameters
‘failures’ Brown [MPa]
mi σ3/σc σ1/σc mb mb/mi s a

15 9 90 15 0.17 0.60 2.07 0.230 0.0242 0.999


Quartzite core 8 9 90 25 0.22 0.72 1.56 0.173 0.0011 0.653
Habimana et al.
(2002) 4 9 90 50 0.17 1.04 4.83 0.537 0.006 0.617

7 9 90 80 0.17 1.08 4.82 0.536 0.0277 0.5

3 points 8 70 5 0.14 0.31 0.823 0.103 0.0031 0.854


digitised

4 points 8 70 15 0.21 0.51 0.804 0.101 0.0021 0.689


Phyllite core digitised
Habimana et al.
(2002) 4 points 8 70 25 0.21 0.56 1.160 0.145 0.0043 0.780
digitised

4 points 8 70 35 0.20 0.57 0.996 0.125 0.0011 0.626


digitised

Limestone core 3 14.9 75 50-59 0.03 0.50 - - - -


Ok Tedi
11 14.9 75 60-69 0.20 1.60 6.28 0.421 0.2714 0.522

14 14.9 75 70-79 0.20 2.00 9.56 0.640 0.6415 0.626

0 - - 80-89 - - - - - -

24 14.9 75 90-100 0.20 2.25 14.93 1.000 1.0000 0.500

Hornfels core 8 39.1 137 50-59 0.09 0.79 5.06 0.129 0.0160 0.500
Ok Tedi
4 39.1 137 60-69 0.00 0.27 - - - -

35 39.1 137 70-79 0.12 1.61 8.51 0.218 0.2181 0.654

45 39.1 137 80-89 0.11 2.87 14.4 0.368 0.3687 0.500

21 39.1 137 90-100 0.12 2.52 39.1 1.000 1.000 0.500

TABLE 12.13 Comparison of proposed relationships between Hoek-Brown criterion parameters and the
GSI with the current recommendations from Hoek et al. (2002).
Current Equations
Proposed Equations
(Hoek et al. 2002)

Equation 12.2

Equation 12.3

Equation 12.4

271
ESTIMATING ROCK MASS PROPERTIES

FIGURE 12.24 Proposed relationship between mb/mi and GSI. The shading represents the values as D varies from 0 to 1.
For D = 1, the proposed and existing relationships are the same.

FIGURE 12.25 Proposed relationship between s and GSI. The shading represents the values as D varies from 0 to 1.
For D = 1, the proposed and existing relationships are the same.

272
12 | Rock Mass Strength

FIGURE 12.26 Proposed relationship between α and GSI.

FIGURE 12.27 Proposed relationship between mb/mi and GSI. and s v GSI for the three broad categories of disturbance
factor, D.

273
ESTIMATING ROCK MASS PROPERTIES

The proposed equations are kept in the same format the proposed equations listed in Table 12.13, which
as those commonly used and as recommended by suggest the following.
Hoek et al. (2002). • For the limestone pillars, GSI = 60 – 75 which gives
• The proposed denominators of the exponents in the parameter a to be approximately 0.5. As s a = 0.11
Equation 12.2 for mb and in Equation 12.3 for s, to 0.17, s can be calculated as equal to 0.012 – 0.029,
differ slightly with an average of 0.021
• The more substantial change is the proposed • or the sandstone pillars,
F
 GSI = 51 – 64 which gives
relationship between the parameter a and GSI. a ≈ 0.5. Again, as s a = 0.11, s = 0.012
Equation 12.4 allows the parameter a to vary • For the iron ore pillars, GSI = 40 – 60, s = 0.01 and
between 0.5 for GSI = 100 and 1.0 for GSI ≤ 10. a = 0.55 (Table 12.12)
Others have also suggested that the parameter a
should vary (Carter et al. 2007; Carter et al. 2008;
• For the large diameter granite core, GSI = 50 – 65,
hence a ≈ 0.5. As s a = 0.04, s = 0.0016.
Diederichs et al. 2007; Mostyn and Douglas 2000;
These values lie within the range shown in Figure
You 2011)
12.25 for the relationship between GSI and s.
• ere is a broad agreement between increasing
Th

disturbance and increasing D and lower mb and s. 12.6.1.1 Ok Tedi Core
The Ok Tedi core data presented in Section 12.5 are
12.6.1 Validation shown as dimensionless principal stress plots in Figure
Two checks of the proposed equations listed in Table 12.28 and Figure 12.29. These plots also show the
12.13 are made. Hoek-Brown strength envelopes predicted using the
The data for the limestone (Esterhuizen et al. 2008), existing GSI relationships and those predicted using
sandstone (Krauland et al. 1989), and iron ore (Hosni the proposed relationships as shown in Table 12.14.
et al. 2015) mine pillars, and the large diameter granite It can be seen in Figure 12.28 and Figure 12.29 that
core Thorpe et al. (1980) were not able to be directly the predicted Hoek-Brown strength envelopes curve-fit
used in the correlations as the rock masses largely the data well. Further, the envelopes predicted using
failed in unconfined compression and hence could not the proposed relationships (Equation 12.2 to Equation
be used to curve-fit unique Hoek-Brown envelopes. 12.4) fit the data better.
Instead the data were used to check the veracity of

TABLE 12.14 Hoek-Brown criterion parameters based on the GSI relationships.


GSI Current Equations Proposed Equations
(Hoek et al. 2002) (Equation 5 14 to Equation 12 4)

mb/mi s a mb/mi s a
55 0.200 0.0067 0.504 0.306 0.0235 0.52

65 0.287 0.0205 0.502 0.398 0.0541 0.51

75 0.409 0.0622 0.501 0.518 0.1245 0.51

85 0.585 0.1889 0.50 0.674 0.2865 0.50

95 0.836 0.5738 0.50 0.877 0.6592 0.50

274
12 | Rock Mass Strength

FIGURE 12.28 Normalised principal stress plots for hornfels (as there is only four UCS data for 60 ≤ GSI ≤ 69, this curve
is not shown.) The curves are the predicted Hoek-Brown envelope using the current GSI relationships (dashed) and those
proposed (solid), for GSI = 55 and 75, respectively.

275
ESTIMATING ROCK MASS PROPERTIES

FIGURE 12.29 Normalised principal stress plots for limestone. The curves are the predicted Hoek-Brown envelope using the
current GSI relationships (dashed) and those proposed (solid), for GSI = 75 and 95, respectively. Note that there was no data
for 80 ≤ GSI ≤ 89. Open symbols are data from limestone breccia.

276
12 | Rock Mass Strength

12.7 CONCLUSIONS
An extensive database of large scale hard rock failures
was assembled from literature. The database comprises
mine pillars in igneous, metamorphic and sedimentary
rocks and both small and large diameter core testing.
In each case the rock mass quality, or data that could
be used to assess rock mass quality, had been recorded.
A distinct contribution is the collation of this extensive
database of large scale rock mass strength and rock
mass quality.
The confining stress range of all but one of the
databases is quite low, typically being between 0 and
0.2σc. In fact, very many of the cases were effectively
unconfined compressive stress failures.
This in itself is a very interesting statistic. Most
large-scale failures experienced in rock masses occur at
either tensile, zero or low confinement. It is clear then
that failure criteria which work in the low confining
stress range are therefore more useful in predicting the
rock mass strength in most experienced cases. 

277
ESTIMATING ROCK MASS PROPERTIES

278
13 | DISL – Coal

13 DISL – COAL

13.1 OUTLINE
Chapter 12 proposed a new set of relationships between in predicting the rock mass strength in most
the Hoek-Brown criterion parameters mb, s and a, and experienced cases.
GSI based on the collation of several databases of rock The one exception is the coal pillar database. The
mass stress failures. The confining stress range of all coal pillar database comprises 162 cases of collapsed
but one of the databases is however quite low, typically pillars ranging up to a confining stress of 2.5σc. This
being between 0 and 0.2σc. database is shown as a normalised principal stress plot
As stated in Chapter 12, this is a very interesting in Figure 13.1. This chapter looks at whether a multi-
statistic. Most of the large-scale failures experienced curve envelope – the damage initiation spalling limit
in rock masses occur at either tensile, zero or low (DISL) approach – as initially proposed by Diederichs
confinement. Failure criteria that work in the low (2000) and as shown in Figure 13.2 – can be fitted to
confining stress range are therefore more useful the coal pillar database.

FIGURE 13.1 Predicted principal stresses normalised with the assumed UCS. Closed symbols are collapsed pillars, open
symbols are un-collapsed pillars. The yielding pillars from Jitpur, India are circled.

279
ESTIMATING ROCK MASS PROPERTIES

13.2 BACKGROUND
As discussed in Chapter 5.2 and Chapter 5.6.2,
Diederichs (2000), Carter et al. (2007) and Diederichs
et al. (2007) working with computer models and test
results of the Lac Du Bonnet granite of AECL’s URL in
Manitoba, Canada, identified an issue with estimating
the Hoek-Brown material parameters for spall-prone
rock masses, which they defined as having mi >> 15
and GSI >> 65. They proposed the relationships in EQUATION 13.1
Equation 13.1 for estimating rock mass strength under
low confining stresses. Diederichs and his co-workers proposed the DISL
Diederichs et al. (2007) noted that the resulting approach to model the strength of spall-prone rock
peak strength parameters ‘will significantly over- masses. The DISL strength envelope is schematically
predict failure if used in non-linear plasticity codes’ shown in Figure 13.2. It comprises four zones as
and hence included the residual values as a second described below in order of increasing confinement.
limit. The change from peak to residual occurring at 1 Under tensile stresses, the rock mass can unravel
approximately a confinement based on the σ1/σ3 ratio 2 Under compressive stresses but below the ‘‘damage
of 7 to 10. threshold’’ the rock is not damaged and remains

FIGURE 13.2 Idealised strength envelope in normalised principal stress, modified from Diederichs et al. (2007)

280
13 | DISL – Coal

undisturbed. Depending upon the confinement


when the damage threshold is exceeded, micro-
crack damage accumulates, leading either to spalling
or macro-scale shear failure. EQUATION 13.5
3 Spalling
Diederichs et al. (2007) proposed an empirically- 5 It is suggested that as confinement continues to
based Hoek-Brown envelope to model this increase, the DISL needs to be extended to include
damage initiation threshold as given in Equation the ‘maximum confined strength’ as advocated by:
13.2. As discussed in Chapter 5.2, systematic „„ Mogi (1966) who made two key observations
cracking starts at compressive stresses between regarding intact rock failure:
0.3σc to 0.5σc, that is UCS*/UCS = 0.3 to 0.5. ww the failure mechanism changes from brittle to
Recalling from Chapter 11.3.7 that for coal mi ductile as the confining stress increases
typically ranges between 14 and 18, with an ww the yield strength (σ1/σ3) becomes constant at
average value of 16 gives the peak strength values: high confining stresses
„„ The idea that the strength of a rock mass is limited
is in keeping with the observations of Bieniawski
(1968) that “there exists a critical pillar size above
which the strength of coal does not change with an
increase in pillar dimensions”
„„ Barton (1976) who coined the term critical
effective confining pressure to be that confining
pressure above which shear strength cannot be
made to increase
„„ Singh et al. (2011) and Singh and Singh (2012)
who suggest that the critical confining pressure
can be assumed to be approximately the UCS
„„ Singh et al. (2015) and Singh et al. (2015) who
suggest that for anisotropic rock masses, the
critical confining pressure is approximately 1.25
EQUATION 13.2 x UCS.
One of the conclusions reached in Chapter 11, is
The residual strength values which are empirically that the maximum confined strength applicable to
defined as s = 0, a = 0.75 and m = mi / 3 give m = 16/3 the coal pillar database is approximately 0.6,
= 5.333 and therefore: with a range between 0.5 and 0.7.


EQUATION 13.3
EQUATION 13.6
4 Macro-scale shear failure
Increasing the confinement changes the controlling Equation 13.2 to Equation 13.5 are overlaid onto the
mechanism to one of macro-scale shear failure. coal pillar database in Figure 13.3 for collapsed pillars
The generalised Hoek-Brown criterion (Hoek et al. and Figure 13.4 for un-collapsed pillars as principal
2002) with parameters provided in in Chapter 11 stress plots. The range for each of the equations is
as reasonable ‘average’ or ‘typical’ values as used by: shaded in the figures with different colours. The best
ww Medhurst and Brown (1998) for mid-bright
fitting set of envelopes for the coal pillar database were
Moura coal (Equation 13.4) visually assessed to comprise:
ww Esterhuizen et al. (2010) for USA coal
„„  amage initiation threshold with UCS*/UCS =
D
(Equation 13.5) 0.3, i.e.

„„ Spalling limit

EQUATION 13.4

281
ESTIMATING ROCK MASS PROPERTIES

The damage initiation threshold and spalling These three best-fitting envelopes are shown
limit capture the USA coal pillars that were recorded superimposed onto the data in Figure 13.5.
as either “massive collapse or rock burst”, which are These three envelopes have then been combined to
shown as the light blue solid diamonds in Figure 13.3). form the extended DISL envelope with the
This supports the view that coal can be thought of as „„ Maximum confined strength, given by
a brittle rock under low confinement and hence the
application of a DISL approach.
„„ Macro-scale shear failure
The extended DISL envelope is then plotted in
Figure 13.6 and Figure 13.7 in principal stress space.
Figure 13.8 and Figure 13.9 show the data in the
deviatoric stress space.

FIGURE 13.3 Normalised average principal stress plots of the collapsed coal pillar database. The light blue diamonds are
“massive collapse or rock burst” USA coal pillars. The right-hand plots are a subset of the data, showing the cases with σ3/σc
≤ 0.5. The ranges for the ‘damage threshold’, ‘spalling limit’ and ‘shear failure’ envelopes are shown.

282
13 | DISL – Coal

FIGURE 13.4 Normalised average principal stress plots of the un-collapsed coal pillar database. The right-hand plots are
a subset of the data, showing the cases with σ3/σc ≤ 0.5. The ranges for the ‘damage threshold’, ‘spalling limit’ and ‘shear
failure’ envelopes are shown.

283
ESTIMATING ROCK MASS PROPERTIES

FIGURE 13.5 Normalised average principal stress plots of the coal pillar database with σ3/σc ≤ 0.5; solid symbols represent
the collapsed pillars (top plots) and open symbols the un-collapsed pillars (bottom plots). The light blue diamonds are
“massive collapse or rock burst” USA coal pillars. The best fitting curves for the ‘damage threshold’, ‘spalling limit’ and ‘shear
failure’ envelopes are shown.shown.

284
13 | DISL – Coal

FIGURE 13.6 Normalised average principal stress plots of the collapsed coal pillar database. The various slender pillars are
encircled. The right-hand plots are a subset of the data, showing the cases with σ3/σc ≤ 0.5. The extended DISL strength
curve is also shown with the MCS equal to 0.5, 0.6 and 0.7.

FIGURE 13.7 Normalised average principal stress plots of the un-collapsed coal pillar database. The yielding pillars from
Jitpur, India are circled as are the highly rectangular pillars from Utah, USA and Australia. The right-hand plots are a subset of
the data, showing the cases with σ3/σc ≤ 0.5. The extended DISL strength curve is also shown with the MCS equal to 0.5, 0.6
and 0.7.

285
ESTIMATING ROCK MASS PROPERTIES

FIGURE 13.8 Normalised average principal deviatoric stress plots of the collapsed coal pillar database. The various slender
pillars. The right-hand plots are a subset of the data, showing the cases with σ3/σc ≤ 0.5. The extended DISL strength curve is
also shown with the MCS equal to 0.5, 0.6 and 0.7.

FIGURE 13.9 Normalised average principal deviatoric stress plots of the un-collapsed coal pillar database. The yielding pillars
from Jitpur, India are circled as are the highly rectangular pillars from Utah, USA and Australia. The right-hand plots are a
subset of the data, showing the cases with σ3/σc ≤ 0.5. The extended DISL strength curve is also shown with the MCS equal
to 0.5, 0.6 and 0.7.

286
13 | DISL – Coal

The extended DISL envelope (Figure 13.7 to Figure • Excluding the Jitpur yielding pillars, ten of the 89
13.9) differentiates between the collapsed and the un- un-collapsed pillars modelled (11.2%) plot above the
collapsed coal pillars very well. This is particularly so envelope: four from the USA, five from Australia,
considering the rather generalised approach taken one from India.
here in assigning UCS and estimating the various
Three of the four USA pillars come from the Utah
parameters. It is interesting to note that the data seems
mines which are identified in Chapter 11.3.3,
to fit a maximum confined strength that is reached at
Figure 11.3 as being highly rectangular. Two of the
ratios of σ3/σc between 0.13 and 0.3. This is much lower
five Australian pillars are also highly rectangular.
than the value of one suggested by Singh et al. (2011).
The details of the cases that do not conform
The extended DISL envelope separates approximately to the envelope are listed in Table 13.1. The
80% of the collapsed and 90% of the un-collapsed highly rectangular pillars and the slender pillars
pillars modelled. are highlighted.
• Twenty of the 96 collapsed pillars modelled (20.8%)
plot below the envelope: nine from South Africa, six Figure 13.10 and Figure 13.11 combine just the
from the USA, five from India. damage initiation threshold with the maximum
Of these 20, 16 are slender pillars, defined as having confined strength. This simpler envelope also does
w/h ≤ 4 (Mark 2006) and of those, nine have a w/h ≤ a very good job, separating the same 80% of the
2, which can be considered to be very slender pillars. collapsed and 90% of the un-collapsed pillars as the
Accepting that slender pillars need to be specifically extended DISL envelope. However, the extended DISL
assessed, only four of the 96 collapsed pillars lie envelope (Figure 13.7 to Figure 13.9) visually fits the
beneath the envelope, i.e. 4.2%. data better.

287
ESTIMATING ROCK MASS PROPERTIES

TABLE 13.1 Details of the coal pillar cases which do not conform to the DISL envelope. Highly
rectangular pillars and the slender pillars are highlighted.
State Mine Seam Depth w/h Pillar σ3/σc σ1/σc
[m] widths
[m]

Collapsed pillars that lie below the DISL envelope

KY 74 326 4.5 22.6 x 16.5 1.27 1.68

KY 75 122 4.6 9.1 x 9.1 0.46 0.80

WV 2 91 1.3 3 x 12.2 0.34 0.54


USA

WV 76 91 2.5 6.1 x 6.1 0.28 0.71

WV 36 69 1.8 6.1 x 6.1 0.14 0.55

WV 23 85 2.0 6.1 x 12.2 0.20 0.66

Begonia Begonia 36 1.3 3.9 x 3.9 0.11 0.44

Amlai Burhar 30 0.8 4.5 x 4.5 0.04 0.30

India Ramnagar Begunia 51 1.7 3x3 0.12 0.52

Kankanee XIII 160 3.0 19.8 x 19.8 0.47 0.90

Pure Kajora Lower 54 1.5 5.4 x 5.4 0.12 0.55


Kajora

Vaal Basin Sigma OFS2b 70 4.3 12.5 0.36 0.52


weak coal

Vaal Basin Sigma OFS3 82 3.6 10.0 0.37 0.65


weak coal

Vaal Basin Sigma OFS2a 104 4 12.0 0.49 0.82


weak coal
South Africa

Vaal Basin Sigma OFS2 108 3.3 10.6 0.47 0.90


weak coal

Vaal Basin Sigma OFS2 112 3.8 10.6 0.52 0.94

Free State Vierfontein Main 60 3.9 7 0.33 0.58

Utrecht Umgala Alfred 51.5 1.5 6 0.18 0.52

Free State Vierfontein Main 21 2.1 7.0 0.09 0.20

Witbank Wolvekrans W2 41 1.0 6.4 0.07 0.42

Un-collapsed pillars that lie above the DISL envelope

KY 100 290 4.6 9.1 x 6.1 1.22 2.26

UT 35 503 3.0 34.4 x 9.1 1.50 3.23


USA

UT 35 594 3.0 34.4 x 9.1 1.76 3.82

UT 3 585 3.9 11.3 x 30.5 1.93 3.48

India Morganpit Salarjung 270 2.7 8.1 x 8.1 0.82 1.80

FS7 510 8.2 23.5 x 44.5 0.56 2.67

FS10 450 9.2 23.5 x 44.5 0.55 2.44

AUSTRALIA FS11 150 4.6 16 x 18 0.23 1.00

FS14 430 10.7 32 x 36 0.40 2.00

FS15 75 5.0 15 x15 0.21 1.03

288
13 | DISL – Coal

FIGURE 13.10 Normalised average principal stress plots of the collapsed coal pillar database. The various slender pillars are
circled. The right-hand plots are a subset of the data, showing the cases with σ3/σc ≤ 0.5. The simplified strength curve is also
shown with MCS equal to 0.6.

FIGURE 13.11 Normalised average principal stress plots of the un-collapsed coal pillar database. The yielding pillars from
Jitpur, India are circled as are the highly rectangular pillars from Utah, USA and Australia. The right-hand plots are a subset of
the data, showing the cases with σ3/σc ≤ 0.5. The simplified strength curve is also shown with MCS equal to 0.6.

289
11 | Coal Pillar Case Studies

13.3 CONCLUSIONS • Damage initiation threshold


The work presented uses the extensive coal pillar
database compiled in Chapter 11 to demonstrate the
applicability of the DISL approach. • Spalling limit
The DISL approach was extended by incorporating
the maximum confined strength. A simplified
envelope was also developed. However, by inspection, • Macro-scale shear failure
the extended DISL envelope provided the better visual
fit to the coal pillar database.
The extended DISL multi-curve envelope provides a
very good predictor for the rock mass strength of coal • Maximum confined strength
pillars. Geological structures need to be considered,
particularly for slender pillars (w/h ≤ 4), but if failure
is expected to occur through the rock mass then the
envelope can be used for the rock mass strength. • In
the absence of better data, a value of σc = 6.5
The Hoek-Brown parameters for the extended MPa can be used to represent the unconfined
multi-curve DISL envelope applicable to coal are compressive strength of a 1 m3 cube of coal
(Figure 13.12):

FIGURE 13.12 Recommended extended DISL envelope for coal pillars (curves defining the shaded area)

290
14 | References

14 REFERENCES

Adhikary, D. P., and Dyskin, A. V. (1997). “A Cosserat continuum Rock Engineering, J. A. Hudson, ed., Pergamon, 155-183.
model for layered materials.” Computers and Geotechnics, 20(1), Bandis, S. C., Lumsden, A. C., and Barton, N. R. (1981).
15-45. “Experimental studies of scale effects on the shear behaviour
Adhikary, D. P., and Dyskin, A. V. (1998). “A continuum model of rock joints.” International Journal of Rock Mechanics, Mining
of layered rock masses with non-associative joint plasticity.” Sciences & Geomechanics Abstracts, 18, 1-21.
International Journal for Numerical and Analytical Methods in Bandis, S. C., Lumsden, A. C., and Barton, N. R. (1983).
Geomechanics, 22, 245-261. “Fundamentals of Rock Joint Deformation.” International
AFTES (2003). “Guidelines for characterisation of rock masses Journal of Rock Mechanics, Mining Sciences & Geomechanics
useful for the design and the construction of underground Abstracts, 20(6), 249-268.
structures.” Barton, N. (1973). “Review of a new shear strength criterion for
Ajalloeian, R., and Mohammadi, M. (2014). “Estimation of rock joints.” Engineering Geology, 7, 287-332.
limestone rock mass deformation modulus using empirical Barton, N. (1976). “The shear strength of rock and rock joints.”
equations.” Bulletin Engineering Geology and the Environment, International Journal of Rock Mechanics, Mining Sciences &
73(2), 541-550. Geomechanics Abstracts, 13, 225-279.
Al-Ajmi, A. M. (2006). “Wellbore stability analysis based on a Barton, N. (1995). “The influence of joint properties in modelling
new true-triaxial failure criterion.”PhD, KTH. jointed rock masses.” Proc., 8th ISRM Congress, 1023-1032.
Al-Ajmi, A. M., and Zimmerman, R. W. (2005). “Relation Barton, N. (1999). “General report concerning some 20th
between the Mogi and the Coulomb failure criteria.” century lessons and the 21st century challenges in applied rock
International Journal of Rock Mechanics and Mining Sciences, mechanics, safety and control of the environment.” Proc., 9th
42(3), 431-439. International Congress on Rock Mechanics.
Alejano, L., Arzua, J., and Perez-Rey, I. (2015). “Effect of scale Barton, N. (2002). “Some new Q-value correlations to assist in
and structure on the strength and deformability of rocks.” site characterisation and tunnel design.” International Journal
Proc., 13th International Symposium on Rock Mechanics - ISRM of Rock Mechanics and Mining Sciences, 39, 185-216.
Congress 2015. Barton, N., and Bandis, S. C. (1980). “Some effects of scale on
Amadei, B. (1996). “Importance of anisotropy when estimating the shear strength of joints.” International Journal of Rock
and measuring insitu stresses in rock.” International Journal Mechanics, Mining Sciences & Geomechanics Abstracts, 17, 69-
of Rock Mechanics, Mining Sciences & Geomechanics Abstracts, 73.
33(3), 293-325. Barton, N., and Bandis, S. C. (1982). “Effects of block size on the
Amadei, B., and Savage, W. Z. (1991). “Analysis of borehole shear behaviour of jointed rock”,76, Issues in Rock Mechanics,
expansion and gallery tests in anisotropic rock masses.” 739-760.
International Journal of Rock Mechanics, Mining Sciences & Barton, N., Lien, R., and Lunde, J. (1974). “Engineering
Geomechanics Abstracts, 28(5), 383-396. classification of rock masses for the design of tunnel support.”
Amadei, B., Savage, W. Z., and Swolfs, H. S. (1987). Rock Mechanics, 6(4), 189-236.
“Gravitational stresses in anisotropic rock masses.” Barton, N., and Pandey, S. K. (2011). “Numerical modelling of
International Journal of Rock Mechanics, Mining Sciences & two stoping methods in two Indian mines using degradation
Geomechanics Abstracts, 24(1), 5-14.
of c and mobilization of ϕ based on Q-parameters.”
Asche, H. R., and Quigley, A. (1999). “Tunnelling Design on International Journal of Rock Mechanics and Mining Sciences,
the Northside Storage Tunnel Project.” Proc., 10th Australian 48(7), 1095-1112.
Tunnelling Conference, AusIMM, 143-154.
Barton, N. R. (1972). “A model study of rock joint deformation.”
AWT (1998). “Upper Canal Inspection.” Unpublished report:, International Journal of Rock Mechanics and Mining Sciences, 9,
Australian Water Technologies, 67. 579-602.
Barton, N. R. (2014). “Shear strength of rock, rock joints and
Badelow, F., Best, R., Bertuzzi, R., and Maconochie, D. (2005). rock masses - Problems and some solutions.” Proc., Rock
“Modelling of defect and rock bolt behaviour in geotechnical engineering and rock mechanics - Structures in and on rock
numerical analysis for the Lane Cove Tunnel.” Proc., masses, Taylor & Francis Group, 3-16.
Geotechnical Aspects of Tunnelling for Infrastructure Projects, Barton, N. R., and Choubey, V. D. (1977). “The shear strength of
AGS-AUCTA Mini Symposium, 1-9. rock joints in theory and practice.” Rock Mechanics, 10, 1-54.
Bahaaddini, M., Sharrock, G., and Hebblewhite, B. K. (2013). Barton, N. R., and Grimstad, E. (2004). “The Q-system
“Numerical investigations of the effect of joint geometrical following 30 years of development and application in
parameters on the mechanical properties of a non-persistent tunneling projects.” Proc., Eurock 2004 & 53rd Geomechanics
jointed rock mass under uniaxial compression.” Computers and Colloquium.
Geosciences, 49, 206-225.
Barton, N. R., and Grimstad, E. (2014). “Forty years with the
Bandis, S. C. (1993). “Engineering Properties and Q-system in Norway and abroad.” Fjellsprengningsteknikk,
Characterization of Rock Discontinuities”,6, Comprehensive Bergmekanikk / Geoteknikk, 4.1-4.25.

291
ESTIMATING ROCK MASS PROPERTIES

Barton, N. R., and Grimstad, E. (2014). “An illustrated guide to Cai, M., and Kaiser, P. K. (2007). “Obtaining modeling
the Q-system ebook (7 May 2015). parameters for engineering design by rock mass
Barton, N. R., and Shen, B. (2016). “Risk of shear failure and characterisation.” Proc., 11th Congress of the International
extensional failure around overstressed excavations in brittle Society for Rock Mechanics, Taylor & Francis Group.
rock.” Journal of Rock Mechanics and Geotechnical Engineering. Cai, M., Kaiser, P. K., Uno, H., Tasaka, Y., and Minami, M.
Battaglia, J., Dixon, S., and Hudson, A. (2006). “Cross city (2004). “Estimation of rock mass deformation modulus and
tunnel, Sydney, NSW.” Proc., ACAA Technical Conference, 14. strength of jointed hard rock masses using the GSI system.”
Benz, T., Schwab, R., Kauther, R. A., and Vermeer, P. A. (2008). International Journal of Rock Mechanics and Mining Sciences,
“A Hoek–Brown criterion with intrinsic material strength 41(1), 3-19.
factorization.” International Journal of Rock Mechanics and Call, R. D., Cicchini, P. F., Ryan, T. M., and Barkley, R. C.
Mining Sciences, 45(2), 210-222. (2000). “Managing and analysing overall pit slopes”,Chapter
Berry, D. S. (1960). “An elastic treatment of ground movement 4, Slope Stability Surface Mining, W. A. Hustrulid, M. K.
due to mining in Isotropic ground.” Journal of the Mechanics McCarter, and D. J. A. van Zyl, eds.
and Physics of Solids, 8(4), 280-292. Carter, T. G., Diederichs, M. S., and Carvallho, J. L. (2007). “A
Bertuzzi, R. (2012). “Discussion on ‘Sari, M., An improved unified procedure for Hoek-Brown prediction of strength and
method of fitting experimental data to the Hoek–Brown post yield behaviour for rockmasses at the extreme ends of the
failure criterion, Engineering Geology (2012)’.” Engineering rock competency scale.” Proc., 11th Congress of the ISRM 2007,
Geology, 152, 213-214. Taylor & Francis Group, 161-164.
Bertuzzi, R. (2014). “Sydney sandstone and shale parameters for Carter, T. G., Diederichs, M. S., and Carvallho, J. L. (2008).
tunnel design.” Australian Geomechanics, 49(1), 1-40. “Application of modified Hoek-Brown transition relationships
Bertuzzi, R. (2014). “Sydney sandstone and shale parameters for for assessing strength and post yield behaviour at both ends
tunnel design - corrigendum 1.” Australian Geomechanics, of the rock competence scale.” Journal of the South African
49(2), 95-104. Institute of Mining and Metallurgy, 108(6), 325-338.
Bertuzzi, R., Douglas, K., and Mostyn, G. (2016). “Comparison Carter, T. G., and Marinos, V. (2014). “Use of GSI for rock
of quantified and chart GSI for four rock masses.” Engineering engineering design.” Proc., Proceedings 1st International
Geology, 202, 24-35. Conference on Applied Empirical Design Methods in Mining,
Bertuzzi, R., Douglas, K., and Mostyn, G. (2016). “Updating the 1-19.
GSI and Hoek-Brown parameters relationships.” International Carvalho, J. L., Carter, T. G., and Diederichs, M. S. (2007). “An
Journal of Rock Mechanics and Mining Sciences, submitted. approach for prediction of strength and post yield behaviour
Bertuzzi, R., and Pells, P. J. N. (2002). “Geotechnical parameters for rock masses of low intact strength.” Proc., 1st Canada-US
of Sydney sandstone and shale.” Australian Geomechanics, Rock Mechanics Symposium, Taylor & Francis, 277-285.
37(5), 41-54. Cecil, O. S. (1970). “Correlations of rock bolt-shotcrete support
Bieniawski, Z. T. (1968). “The effect of specimen size on and rock quality parameters in Scandinavian tunnels.” PhD,
compressive strength of coal.” International Journal of Rock University of Illinois, Urbana, USA.
Mechanics and Mining Sciences, 5, 325-335. Christensen, R. M. (1997). “Yield functions/failure criteria
Bieniawski, Z. T. (1968). “In situ strength and deformation for isotropic materials.” Proceedings of the Royal Society A:
characteristics of coal.” Engineering Geology, 2(5), 325-340. Mathematical, Physical and Engineering Sciences, 453(1962),
Bieniawski, Z. T. (1973). “Engineering classification of jointed 1473-1491.
rock masses.” Transactions South African Institution of Civil Christensen, R. M. (2004). “A two property yield, failure
Engineers, 15(2), 335-342. (fracture) criterion for homogenous, isotropic materials.”
Bieniawski, Z. T. (1974). “Estimating the strength of rock Journal of Engineering Materials and Technology, 126, 45-52.
materials.” Journal of the South African Institute of Mining and Christensen, R. M. (2005). “Exploration of ductile, brittle
Metallurgy, 74(8), 312-320. failure characteristics through a two parameter yield/ failure
Bieniawski, Z. T. (1976). “Rock mass classifications in rock criterion.” Material Science and Engineering, A394, 417-424.
engineering.” Proc., Proceedings of the Symposium on Exploration Christensen, R. M. (2006). “A comparative evaluation of three
for Rock Engineering, A.A. Balkema. isotropic, two property failure theories.” Journal of Applied
Bieniawski, Z. T. (1978). “Determining rock mass deformability: Mechanics, 73, 852-859.
experience from case histories.” International Journal of Rock Christensen, R. M. (2006). “Yield functions and plastic potentials
Mechanics, Mining Sciences & Geomechanics Abstracts, 15, 237- for BCC Metals and possibly other materials.” Journal of
247. Mechanics of Materials and Structures, 1, 195-212.
Bieniawski, Z. T. (1989). Engineering Rock Mass Classifications, Christensen, R. M. (2007). “A Comprehensive theory of yielding
John Wiley & Sons, New York. and failure for isotropic materials.” Journal of Engineering
Brady, B. H. G., and Brown, E. T. (2005). Rock Mechanics for Materials and Technology, 129(2), 173.
underground mining, Kluwer Academic Publishers. Chun, B.-S., Ryu, W. R., Sagong, M., and Do, J.-N. (2009).
Branagan, D. F. (2000). “The Hawkesbury Sandstone: its orgins “Indirect estimation of the rock deformation modulus based
and later life.” Proc., Sandstone City - Sydney’s dimension on polynomial and multiple regression analyses of the RMR
stone and other sandstone geomaterials, Geological Society of system.” International Journal of Rock Mechanics and Mining
Australia, 23-38. Sciences, 46(3), 649-658.
Branagan, D. F., and Packham, G. H. (2000). Field Geology of Clark, P. (2005). “Monitoring in MC1A and MC2A between
New South Wales, NSW Department of Mineral Resources, chainages 1300 and 1600 and a possible reason for the
Sydney. shotcrete cracking.” Unpublished report:, TJH Joint Venture,
Bristow, J. R. (1960). “Microcracks and the static and dynamic Lane Cove Tunnel Project.
constants of annealed and heavily cold worked metals.” British Clarke, S., and Pells, P. J. N. (2004). “A large scale cable jacking
Journal Applied Physics, 11, 81-85. test for rock mass modulus measurement, Lucas Heights,
Brown, E. T. (2008). “Estimating the mechanical properties of Sydney.” Proc., Proceedings of the 9th Australia New Zealand
rock masses.” Proc., SHIRMS 2008, Australian Centre for Conference on Geomechanics, AGS, 152-158.
Geomechanics, 3-21. Clarke, S. J., de Ambrosis, A., Bertuzzi, R., and Redelinghuys,
Brown, E. T., and Hoek, E. (1988). “Discussion on paper by Ucar J. (2014). “Design and construction for the widening of the
“Determination of shear failure envelope in rock masses”.” M2 Norfolk twin tunnels.” Proc., 15th Australasian Tunnelling
Journal of Geotechnical Engineering, 114, 371-373. Conference, Australian Tunnelling Society.

292
14 | References

Colmenares, L. B., and Zoback, M. D. (2002). “A statistical University.


evaluation of intact rock failure criteria constrained by Duncan Fama, M. E., Trueman, R., and Craig, M. S. (1995).
polyaxial test data for five different rocks.” International “Two- and three- dimensional elasto-plastic analysis for
Journal of Rock Mechanics and Mining Sciences, 39, 695-729. coal pillar design and its application to highwall mining.”
Colwell, M. G. (2010). “Pillar design procedures and research International Journal of Rock Mechanics, Mining Sciences &
methodologies - can there or should there be a unified Geomechanics Abstracts, 32(3), 215-225.
approach?” Proc., Second Australian Ground Control in Mining Duncan, J. M., and Goodman, R. E. (1968). Finite Element
Conference, AusIMM, 67-77. Analyses of Slopes in Jointed Rock: A Report of an Investigation,
Cook, A. B. (1994). “Coal and progress: the South African Defense Technical Information Center, University of
story.” The Journal of The South African lnstitute of Mining and California, Berkeley. College of Engineering. Office of
Metallurgy, November / December 1994, 329-335. Research Services.
Costamagna, R., and Bruhns, O. T. (2007). “A four-parameter
criterion for failure of geomaterials.” Engineering Structures, Eberhardt, E. (2012). “The Hoek–Brown Failure Criterion.” Rock
29(3), 461-468. Mechanics and Rock Engineering, 45(6), 981-988.
Cundall, P. A., Pierce, M. E., and Mas Ivars, D. (2008). Edelbro, C. (2003). “Rock mass strength: a review.” Lulea
“Quantifying the size effect of rock mass strength.” Proc., 1st University of Technology, Lulea, Sweden, 160.
Southern Hemisphere International Rock Mechanics Symposium, Edelbro, C., Sjöberg, J., and Nordlund, E. (2007). “A quantitative
Australian Centre for Geomechanics, 3-15. comparison of strength criteria for hard rock masses.”
Tunnelling and Underground Space Technology, 22(1), 57-68.
Das, M. N. (1986). “Influence of width/height ratio on post- Enever, J. R. (1999). “Near surface in-situ stress and its
failure behaviour of coal.” International Journal of Mining and counterpart at depth in the Sydney metropolitan area.”
Geological Engineering, 4(1), 79-87. Australian Geomechanics (June 1999), 65-76.
David, E. C., and Zimmerman, R. W. (2011). “Elastic moduli Esterhuizen, G. S., Dolinar, D. R., and Ellenberger, J. L. (2008).
of solids containing spheroidal pores.” International Journal of “Pillar strength and design methodology for stone mines.”
Engineering Science, 49(7), 544-560. Proc., 27th International Conference on Ground Control in
de Ambrosis, L. P., and Kotze, G. P. (2004). “Stress induced roof Mining, West Virginia University, 241-253.
collapses during construction of the Sydney LPG storage Esterhuizen, G. S., Dolinar, D. R., and Ellenberger, J. L. (2011).
cavern.” Proc., 9th Australian New Zealand Conference on “Pillar strength in underground stone mines in the United
Geomechanics, 159-165. States.” International Journal of Rock Mechanics and Mining
Deere, D. U. (1968). “Geological considerations”,1, Rock Sciences, 48(1), 42-50.
Mechanics in Engineering Practice, K. C. Stagg, and O. C. Esterhuizen, G. S., Iannacchione, A. T., Ellenberger, J. L., and
Zienkiewicz, eds., John Wiley & Sons, 1-20. Dolinar, D. R. (2006). “Pillar stability issues based on a survey
Descamps, F., and Tshibangu, J. P. (2007). “Modelling the of pillar performance in underground limestone mines.” Proc.,
Limiting Envelopes of Rocks in the Octahedral Plane.” Oil & 25th International Conference on Ground Control in Mining,
Gas Science and Technology - Revue de l’IFP. West Virginia University.
Diederichs, M. S. (2000). “Instability of hard rockmasses: The Esterhuizen, G. S., Mark, C., and Murphy, M. M. (2010). “The
role of tensile damage and relaxation.” PhD, University of ground response curve, pillar loading and pillar faiulure in coal
Waterloo, Waterloo, Ontario, Canada. mines.” Proc., 29th International Conference on Ground Control
Diederichs, M. S. (2003). “Rock fracture and collapse under in Mining, 19-27.
low confinement conditions.” Rock Mechanics and Rock Esterhuizen, G. S., Mark, C., and Murphy, M. M. (2010).
Engineering, 36(5), 339-381. “Numerical model calibration for simulating coal pillars, gob
Diederichs, M. S. (2007). “The 2003 Canadian Geotechnical and overburden response.” Proc., 29th International Conference
Colloquium: Mechanistic interpretation and practical on Ground Control in Mining, 46-57.
application of damage and spalling prediction criteria for deep Ewy, R. T. (1999). “Wellbore stability predictions by use of a
tunnelling.” Canadian Geotechnical Journal, 44(9), 1082-1116. modified Lade criterion.” SPE Drill & Completion, 14(2), 85-
Diederichs, M. S., Carvalho, J. L., and Carter, T. G. (2007). 91.
“A modified approach for prediction of strength and post
yield behaviour for high GSI rock masses in strong, brittle Fairhurst, C. (1971). “Fundamental considerations relating to
ground.” Proc., Rock Mechanics: Meeting Society’s Challenges and the strength of rock.” Colloquium on Rock Fracture, Revised
Demands, Taylor & Francis Group, 249-257. from original published in Veröff. Inst. Bodenmechanik und
Diederichs, M. S., Kaiser, P. K., and Eberhardt, E. (2004). Felsmechanik (Karlsruhe), 55, 1-56, 1971.
“Damage initiation and propagation in hard rock during Flores, G., and Karzulovic, A. (2002). “Geotechnical guidelines
tunnelling and the influence of near-face stress rotation.” for a transition from open pit to underground mining.”
International Journal of Rock Mechanics and Mining Sciences, Prepared for international caving study, Stage II, JKMRC,
41(5), 785-812. Brisbane.
Dinc, O. S., Sonmez, H., Tunusluoglu, C., and Kasapoglu, Fookes, P. G. (1997). “Geology for engineers: the geological
K. E. (2011). “A new general empirical approach for the model, prediction and performance.” Quarterly Journal of
prediction of rock mass strengths of soft to hard rock masses.” Engineering Geology, 30, 293-424.
International Journal of Rock Mechanics and Mining Sciences,
48(4), 650-665. Gale, W. (1998). “Coal pillar design issues in longwall mining.”
Douglas, K. J. (2002). “The shear strength of rock masses.” PhD, Proc., Coal Operators’ Conference, University of Wollongong &
University of New South Wales, Sydney, Australia. the AusIMM, 133-146.
Douglas, K. J., and Mostyn, G. (1999). “Strength of large rock Galera, J. M., Alvarez, M., and Bieniawski, Z. T. (2005).
masses – field verification.” Proc., Proceedings of the American “Evaluation of the deformation modulus of rock masses:
Rock Mechanics Association, Balkema, 271-276. comparison of pressuremeter and dilatometer tests with RMR
Drucker, D. C. and Prager, W. (1952). Soil mechanics and plastic prediction.” Proc., ISP5 - Pressio International Symposium.
analysis for limit design. Quarterly of Applied Mathematics, vol. Galvin, J. M. (1995). “Roadway and pillar mechanics workshop.”
10, no. 2, pp. 157–165. Strata Control for Coal Mine Design, Key Centre for Mines,
Du, X., Lu, J., Morsy, K., and Peng, S. (2008). “Coal pillar University of New South Waes.
design formulae review and analysis.” Proc., 27th International Galvin, J. M. (2006). “Considerations associated with the
Conference on Ground Control in Mining, West Virginia

293
ESTIMATING ROCK MASS PROPERTIES

application of the UNSW and other pillar design formulae.” theory.” International Journal of Rock Mechanics and Mining
Proc., 41st US Symposium on Rock Mechanics, ARMA / USRMS. Sciences, 47(4), 568-582.
Galvin, J. M., Hebblewhite, B. K., and Salamon, M. D. G. Hardy, M. P., Hudson, J. A., and Fairhurst, C. (1973). “The
(1999). “UNSW pillar strength determinations for Australian failure of rock beams: Part I - theoretical studies.” International
and South African conditions.” Proc., 37th US Rock Mechanics Journal of Rock Mechanics and Mining Sciences, 10, 53-67.
Symposium, NIOSH, 63-71. Hawkins, A. B. (1998). “Aspects of rock strength.” Bulletin
Gee, R. A., Parker, C. J., and Cuttler, R. J. (2002). “Northside Engineering Geology and the Environment, 57, 17-30.
Storage Tunnel, Sydney: investigation, design & construction.” He, P.-f., Kulatilake, P. H. S. W., Liu, D.-q., and He, M.-c.
Proc., ITA General Congress 2002, IEAust, 149-157. (2016). “Development of new three-dimensional coal mass
Geoderis (2008). “Field trip in the Lorraine iron ore basin.” Proc., strength criterion.” International Journal of Geomechanics.
International Symposium Post-Mining Université J. Fourier. Hencher, S. R., and Richards, L. R. (1989). “Laboratory direct
Gercek, H. (2007). “Poisson’s ratio values for rocks.” International shear testing of rock discontinuities.” Ground Engineering, 24-
Journal of Rock Mechanics and Mining Sciences, 44, 1-13. 31.
Gerrard, C. M. (1982). “Elastic models of rock masses having Hencher, S. R., and Richards, L. R. (2014). “Assessing the Shear
one, two and three sets of joints.” International Journal of Rock Strength of Rock Discontinuities at Laboratory and Field
Mechanics, Mining Sciences & Geomechanics Abstracts, 19, 15- Scales.” Rock Mechanics and Rock Engineering.
23. Hencher, S. R., Toy, J. P., and Lumsden, A. C. (1993). “Scale
Goodman, R. E. (1968). “Effect of joints on the strength of dependent shear strength of rock joints.” Proc., Second
tunnels.” Omaha District, Corps of Engineers. International workshop on scale effects in rock masses, Balkema,
Goodman, R. E. (1974). “The mechanical properties of joints.” 233-240.
Proc., 3rd Congress of the International Society of Rock Mechancs, Hoek, E. (1965). “Rock fracture under static stress conditions.”
127-140. PhD, University of Cape Town.
Goodman, R. E., Taylor, R. L., and Brekke, T. L. (1968). “A Hoek, E. (1983). “The strength of jointed rock masses.”
model for the mechanics of jointed rock.” Journal of the Soil Geotechnique, 23(3), 187-223.
Mechanics and Foundations Division, ASCE, 94(No. SM3, Hoek, E. (1994). “Strength of rock and rock masses.” ISRM News
Proc. Paper 5937), 637-659. Journal 2(2), 4-16.
Goricki, A., Rachaniotis, N., Hoek, E., Marinos, P., Tsotsos, Hoek, E. (1999). “Putting numbers to geology - an engineer’s
S., and Schubert, W. (2006). “Support decision criteria for viewpoint.” Quarterly Journal of Engineering Geology, 32, 1-19.
tunnels in fault zones.” Felsbau, 51-57. Hoek, E. (2001). “Rock mass properties for underground mines”,
Grechka, V., and Kachanov, M. (2006). “Effective elasticity of Underground Mining Methods: Engineering Fundamentals and
fractured rocks: a snapshot of the work in progress.” Geophysics, International Case Studies, W. A. Hustrulid, and R. L. Bullock,
71(6), W45-W58. eds., Society for Mining, Metallurgy, and Exploration,
Grechka, V., and Kachanov, M. (2006). “Effective elasticity of Litleton, Colorado.
rocks with closely spaced and intersecting cracks.” Geophysics, Hoek, E. (2002). “A brief history of the development of the
71(3), D85-D91. Hoek-Brown failure criterion.”RocScience website.
Grgic, D., Homand, F., and Dagallier, G. (2001). “Ageing Hoek, E. (2007). “Rock mass properties”,12, Practical Rock
of Lorraine (France) abandoned iron mines.” Proc., Rock Engineering, https://www.rocscience.com/learning/hoek-s-
Mechanics - a Challenge for Society: Proceedings of the ISRM corner/books.
Regional Symposium Eurock 2001, A A Balkema, 825-830. Hoek, E. (2012). “Blast damage factor D.” RocNews.
Grgic, D., Homand, F., and Hoxha, D. (2003). “A short- and Hoek, E., and Bieniawski, Z. T. (1965). “Brittle rock fracture
long-term rheological model to understand the collapses of propagation in rock under compression.” International Journal
iron ore mines in Lorraine, France.” Computers and Geosciences, of Fracture Mechanics, 1(3), 137-155.
30, 557-570. Hoek, E., and Brown, E. T. (1980). “Empirical strength and
Griffith, A. A. (1921). “The phenomena of rupture and flow in criterion for rock masses.” Journal of Geotechnical Engineering,
solids.” Philosophical Transactions of the Royal Society of London. 106(GT9), 1013-1035.
Series A, Containing papers of a mathematical or physical Hoek, E., and Brown, E. T. (1980). Underground excavations in
character, 221, 163-198. rock, The Institution of Mining and Metallurgy, London.
Grimstad, E. (2007). “The Norwegian method of tunnelling – a Hoek, E., and Brown, E. T. (1997). “Practical estimates of rock
challenge for support design.” Proc., XIV European Conference mass strength.” International Journal of Rock Mechanics and
on Soil Machanics and Geotechnical Engineering. Mining Sciences, 34(8), 1165-1186.
Grimstad, E., and Barton, N. (1993). “Updating of the Q-System Hoek, E., Carranza-Torres, C., and Corkum, B. (2002). “Hoek-
for NMT.” Proc., Proceedings of the International Symposium Brown failure criterion - 2002 edition.” Proc., 5th North
on Sprayed Concrete - Modern use of wet mix sprayed concrete for American Rock Mechanics Symposium, University of Toronto
underground support, Norwegian Concrete Association, 46-66. Press, 267-273.
Hoek, E., Carter, T. G., and Diederichs, M. S. (2013).
Habimana, J., Labiouse, V., and Descoeudres, F. (2002). “Quantification of the Geological Strength Index chart.” Proc.,
“Geomechanical characterisation of cataclastic rocks: 47th US Rock Mechanics / Geomechanics Symposium, ARMA
experience from the Cleuson-Dixence project.” International 13-672.
Journal of Rock Mechanics and Mining Sciences, 39, 677-693. Hoek, E., and Diederichs, M. S. (2006). “Empirical estimation of
Hajiabdolmajid, V., Kaiser, P. K., and Martin, C. D. (2002). rock mass modulus.” International Journal of Rock Mechanics
“Modelling brittle failure of rock.” International Journal of Rock and Mining Sciences, 43(2), 203-215.
Mechanics and Mining Sciences, 39, 731-741. Hoek, E., and Marinos, P. (2007). “A brief history of the
Halakatevakis, N., and Sofianos, A. I. (2010). “Correlation of development of the Hoek-Brown failure criterion.” Soils and
the Hoek–Brown failure criterion for a sparsely jointed rock Rocks, 2.
mass with an extended plane of weakness theory.” International Hoek, E., Marinos, P., and Benissi, M. (1998). “Applicability
Journal of Rock Mechanics and Mining Sciences, 47(7), 1166- of the geological strength index (GSI) classification for very
1179. weak and sheared rock masses. The case of the Athens Schist
Halakatevakis, N., and Sofianos, A. I. (2010). “Strength of a Formation.” Bulletin Engineering Geology and the Environment,
blocky rock mass based on an extended plane of weakness 57, 151-160.

294
14 | References

Hoek, E., Marinos, P. G., and Marinos, V. P. (2005). Kayabasi, A., Gokceoglu, C., and Ercanoglu, M. (2003).
“Characterisation and engineering properties of tectonically “Estimating the deformation modulus of rock masses: a
undisturbed but lithologically varied sedimentary rock comparative study.” International Journal of Rock Mechanics
masses.” International Journal of Rock Mechanics and Mining and Mining Sciences, 40, 55-63.
Sciences, 42, 277-285. Kim, M. K., and Lade, P. V. (1984). “Modelling rock strength in
Hoek, E., and Martin, C. D. (2014). “Fracture initiation and three dimensions.” International Journal of Rock Mechanics,
propagation in intact rock – A review.” Journal of Rock Mining Sciences & Geomechanics Abstracts, 21(1), 21-33.
Mechanics and Geotechnical Engineering, 6(4), 287-300. Krahn, J., and Morgenstern, N. R. (1979). “The ultimate
Holland, C. T. (1964). “The strength of coal in mine pillars.” frictional resistance of rock discontinuities.” International
Proc., 6th U.S Symposium on Rock Mechanics, American Rock Journal of Rock Mechanics, Mining Sciences & Geomechanics
Mechanics Association. Abstracts, 16, 127-133.
Hosni, A., Hadadou, R., Josien, J. P., Piguet, J. P., and Baroudi, Krauland, N., Soder, P., and Agmalm, G. (1989). “Determination
H. (2015). “Subsidence kinetics above room and pillar mines of rock mass strength by rock mass classification - some
& development of kinetic criterion based on retro-analysis of experiences and questions from Boliden Mines.” International
subsidence cases.” Proc., 13th International Symposium on Rock Journal of Rock Mechanics, Mining Sciences and Geomechanical
Mechanics - ISRM Congress 2015. Abstracts, 26(1), 115-123.
Hu, Y., and McMechan, G. A. (2009). “Comparison of effective Kulhawy, F. H. (1975). “Stress deformation properties of rock and
stiffness and compliance for characterising cracked rocks.” rock discontinuities.” Engineering Geology, 9, 327-350.
Geophysics, 74(2), D49-D55. Kulhawy, F. H. (1978). “Geomechanical model for rock
Hudson, J. A. (1980). “Overall properties of a cracked solid.” foundation settlement.” Journal of Geotechnical Engineering,
Mathematical Proceedings of the Cambridge Philosophical GT2, 211-227.
Society, 88(02), 371-384. Kulhawy, F. H., and Mayne, F. H. (1990). “Manual on estimating
Hudson, J. A., and Harrison, J. P. (1997). “An introduction to soil properties for foundation design, EL-6800, Research
the principles”,Chapter 1.1, Engineering Rock Mechanics, Project 1493-6.” Cornell University.
Pergamon, Oxford, England. Kveldsvik, V., Nilsen, B., Einstein, H. H., and Nadim, F. (2007).
3
“Alternative approaches for analyses of a 100,000 m rock slide
Idris, M. A., Saiang, D., and Nordlund, E. (2015). “Stochastic based on Barton–Bandis shear strength criterion.” Landslides,
assessment of pillar stability at Laisvall mine using Artificial 5(2), 161-176.
Neural Network.” Tunnelling and Underground Space Kwasniewski, M. (2013). “Mechanical behaviour of rocks under
Technology, 49, 307-319. true triaxial compression conditions - A review.” Proc., True
Ismael, M. A., Imam, H. F., and El-Shayeb, Y. (2014). “A Triaxial Testing of Rocks, CRC Press, Leiden, Germany, 99-138.
simplified approach to directly consider intact rock anisotropy
in Hoek-Brown failure criterion.” Journal of Rock Mechanics Lade, P. V. (1977). “Elasto-plastic stress-strain theory for
and Geotechnical Engineering, 6(5), 486-492. cohesionless soil with curved yield surfaces.” International
ISRM (1978). “Suggested methods for the quantitative Journal of Solids and Structures, 13, 1019-1035.
description of discontinuities in rock masses.” International Lade, P. V. (1982). “Three-parameter failure criterion for
Society for Rock Mechanics, International Journal of Rock concrete.” Journal of Engineering Mechanics, 108(EM5), 850-
Mechanics, Mining Sciences and Geomechanical Abstracts, 319- 863.
368. Lade, P. V. (1984). “Failure criterion for frictional materials”,20,
Itasca (2008). “FLAC 6.0 Manual.” Itasca International. Mechanics of Engineering Materials, C. S. Desai, and R. H.
Gallagher, eds., John Wiley & Sons, 385-402.
Jaeger, J. C. (1966). “Brittle fracture of rocks”,1, Failure and Lade, P. V. (1993). “Rock strength criteria: The theories and the
Breakage of Rocks, 3-57. evidence”,11, Comprehensive Rock Engineering, J. A. Hudson,
Jaeger, J. C., and Cook, N. G. (1979). Fundamentals of Rock ed., Pergamon, 255-284.
Mechanics, 3rd Edition, Chapman & Hall, London. Lade, P. V. (1997). “Modelling the strengths of engineering
Jaiswal, A., and Shrivastva, B. K. (2009). “Numerical simulation materials in three dimensions.” Mechanics of Cohesive-Frictional
of coal pillar strength.” International Journal of Rock Mechanics Materials, 2, 339-356.
and Mining Sciences, 46, 779-788. Lade, P. V. (2013). “Estimating the parameters for 3D failure
Jang, H.-S., and Jang, B.-A. (2014). “New Method for Shear criterion.” Proc., True Triaxial Testing of Rocks, CRC Press.
Strength Determination of Unfilled, Unweathered Rock Joint.” Lade, P. V., and Duncan, J. M. (1973). “Cubical triaxial tests on
Rock Mechanics and Rock Engineering, 48(4), 1515-1534. cohesionless soils.” Journal of Soil Mechanics and Foundations
Jawed, M., Sinha, R. K., and Sengupta, S. (2013). “Chronological Division, SM10, 793-812.
development in coal pillar design for bord and pillar workings: Lade, P. V., and Duncan, J. M. (1975). “Elastoplastic stress-
A critical appraisal.” Journal of Geology and Mining Research, strain theory for cohesionless soils.” Journal of Geotechnical
5(1), 1-11. Engineering, GT10(107), 1037-1053.
Jiang, H., Wang, X., and Xie, Y. (2011). “New strength criteria Lade, P. V., and Kim, M. K. (1995). “Single hardening
for rocks under polyaxial compression.” Canadian Geotechnical constitutive model for soil, rock and concrete.” International
Journal, 48(8), 1233-1245. Journal of Solids and Structures, 32(14), 1963-1978.
Johnston, I. W., Siedel, J. P., and Haberfield, C. M. (1993). Lan, H., Martin, C. D., and Hu, B. (2010). “Effect of
“Cyclic constant normal stiffness direct shear testing device for heterogeneity of brittle rock on micromechanical extensile
soft rock.” Proc., International Symposium on Hard Soils - Soft behavior during compression loading.” Journal of Geophysical
Rocks, 997-984. Research, 115(B1).
Lee, Y.-K., Pietruszczak, S., and Choi, B.-H. (2012). “Failure
Kaiser, P. K., Diederichs, M. S., Martin, C. D., Sharp, J., and criteria for rocks based on smooth approximations to Mohr–
Steiner, W. (2000). “Underground works in hard rock Coulomb and Hoek–Brown failure functions.” International
tunnelling and mining.” Proc., GeoEng 2000, AGS, 19-24. Journal of Rock Mechanics and Mining Sciences, 56, 146-160.
Kallu, R. R., Keffeler, E. R., Watters, R. J., and Agharazi, A. Lisjak, A., and Grasselli, G. (2014). “A review of discrete
(2015). “Development of a multivariate empirical model for modeling techniques for fracturing processes in discontinuous
predicting weak rock mass modulus.” International Journal of rock masses.” Journal of Rock Mechanics and Geotechnical
Mining Science and Technology. Engineering.

295
ESTIMATING ROCK MASS PROPERTIES

Lunder, P. J. (1994). “Hard rock pillar strength estimation: An properties of South African coals.” Proc., 13th International
applied empirical approach.” Master of Applied Science MASc, Symposium on Rock Mechanics - ISRM Congress 2015.
British Columbia. Maybee, W. G. (2000). “Pillar design in hard brittle rocks.”
MAppSc, Laurentian University, Sudbury, Ontario, Canada.
MacGregor, J. P. (1985). “Wianamatta Group - distribution, McClintock, F. A., and Walsh, J. B. (1962). “Friction on Griffith
stratigraphy and structure”, Engineering Geology of the Sydney cracks in rocks under pressure.” Proc., 4th US National
Region, P. J. N. Pells, ed., AA Balkema, Rotterdam, 139-155. Congress of Applied Rock Mechanics Proceedings, 1015-1021.
Marinos, P., and Hoek, E. (2000). “GSI: a geologically friendly McMahon, B. K. (1983). “Some practical considerations for
tool for rock mass strength estimation.” Proc., ISRM estimating shear strength of joints.” Proc., International
International Symposium, International Society for Rock Symposium on Fundamentals of Rock Joints, 475-485.
Mechanics. McMahon, B. K. (1985). “Geotechnical design in the face of
Marinos, P., and Hoek, E. (2001). “Estimating the geotechnical uncertainty.” Australian Geomechanics Society (E H Davis
properties of heterogeneous rock masses such as Flysch.” Memorial Lecture).
Bulletin Engineering Geology and the Environment, 60, 85-92. McNally, G. (2000). “Introduction - sandstone and the Sydney
Marinos, P., Hoek, E., and Marinos, V. (2006). “Variability of environment.” Proc., Sandstone City - Sydney’s dimension
the engineering properties of rock masses quantified by the stone and other sandstone geomaterials, Geological Society of
geological strength index: the case of ophiolites with special Australia, ix-xiv.
emphasis on tunnelling.” Bulletin of Engineering Geology and McQueen, L. B. (2000). “Stress relief effects in sandstones in
the Environment, 65(2), 129-142. Sydney undergound and deep excavations.” Proc., Sandstone
Marinos, P. V. (2010). “New proposed GSI classification charts City - Sydney’s dimension stone and other sandstone geomaterials,
for weak or for complex rock masses.” Proc., 12th International Geological Society of Australia, 309-329.
Congress, Bulletin of the Geological Society of Greece, 1248-1258. McQueen, L. B. (2004). “In situ rock stress and its effect
Marinos, V. (2014). “Tunnel behaviour and support associated in tunnels and deep excavations in Sydney.” Australian
with the weak rock masses of flysch.” Journal of Rock Mechanics Geomechanics, 39(3), 43-58.
and Geotechnical Engineering. Medhurst, T. P., and E. T. Brown. ‘A Study of the Mechanical
Mark, C. (1999). “The start of the art in coal pillar design.” Society Behaviour of Coal for Pillar Design’. International Journal of
of Mining Metallurgy and Exploration, 308, 1-8. Rock Mechanics and Mining Sciences 35, no. 8 (1998): 1087–
Mark, C. (2006). “The evolution of intelligent coal pillar design: 1105.
1981-2006.” Proc., 25th International Conference on Ground Melkoumian, N., Priest, S. D., and Hunt, S. P. (2009). “Further
Control in Mining, 325-334. Development of the Three-Dimensional Hoek–Brown Yield
Mark, C., and Barton, T. M. (1997). “Pillar design and coal Criterion.” Rock Mechanics and Rock Engineering, 42(6), 835-
strength.” Proc., Proceedings: New technology for ground control 847.
in retreat mining, NIOSH, 49-59. Mishra, A. (2009). “Assessment of coal quality of some Indian
Mark, C. and Chase, F.E. (1997). “Analysis of Retreat Mining coals.” Bachelor, National Institute of Technology, Rourkela.
Pillar Stability.” Proceedings of the New Technology for Ground Mogi, K. (1966). “Pressure dependence of rock strength and
Control in Retreat Mining, NIOSH Publication No. 97-122, transition from brittle fracture to ductile flow.” Bulletin of the
IC9446, pp. 17-34. Earthquake Research Institute, 44, 215-232.
Mark, C., and Gadde, M. (2010). “Global trends in coal mine Mogi, K. (1967). “Effect of the intermediate principal stress on
horizontal stress measurements.” Proc., Coal 2010: Coal rock failure.” Journal of Geophysical Research, 72(20), 5117-
Operators’ Conference, University of Wollongong & AusIMM, 5131.
21-39. Mogi, K. (1971). “Fracture and flow of rocks under high triaxial
Martin, C. D. (1990). “Characterizing in situ stress domains at compression.” Journal of Geophysical Research, 76, 1255-1269.
the AECL URL.” Canadian Geotechnical Journal, 27, 631-646. Mogi, K. (1972). “Fracture and flow of rocks.” Tectonophysics,
Martin, C. D. (1993). “The strength of massive Lac du Bonnet 541-568.
granite around underground opening.” Doctor of Philosophy Mohammadi, H., and Rahmannejad, R. (2010). “The estimation
PhD, Manitoba. of rock mass deformation modulus using regression and
Martin, C. D., and Chandler, N. A. (1994). “The progressive artificial neural networks analysis.” The Arabian Journal for
failure of Lac du Bonnet granite.” International Journal of Rock Science and Engineering, 35(1A), 205-217.
Mechanics, Mining Sciences & Geomechanics Abstracts, 31(6), Mooney, D. W. (2017). “Seismic Waves. An introduction.”
643-659. <https://escweb.wr.usgs.gov/share/mooney/Caribbean3.ppt>.
Martin, C. D., Kaiser, P. K., and McCreath, D. R. (1999). “Hoek- (18 January 2017).
Brown parameters for predicting the depth of brittle failure Mortara, G. (2010). “A yield criterion for isotropic and cross-
around tunnels.” Canadian Geotechnical Journal, 36, 136-151. anisotropic cohesive-frictional materials.” International Journal
Martin, C. D., and Maybee, W. G. (2000). “The strength of for Numerical and Analytical Methods in Geomechanics, 34,
hard-rock pillars.” International Journal of Rock Mechanics and 953-977.
Mining Sciences, 37, 1239-1246. Mostyn, G. R., and Douglas, K. J. (2000). “Strength of intact
Martin, C. D., Read, R. S., and Martino, J. B. (1997). rock and rock masses.” Proc., GeoEng2000, Technomic,
“Observations of brittle failure around a circular test tunnel.” Lancaster, 1389-1421.
International Journal of Rock Mechanics and Mining Sciences, Mostyn, G. R., Helgstedt, M. D., and Douglas, K. J. (1997).
34(7), 1065-1073. “Towards field bounds on rock mass failure criteria.”
Mas Ivars, D. (2010). “Bonded particle model for jointed rock International Journal of Rock Mechanics and Mining Sciences,
mass.” Doctoral Thesis, Royal Institute of Technology (KTH). 34(3-4).
Masoumi, H. (2013). “Investigation into the mechanical Murrell, S. A. F. (1963). “A criterion for brittle fracture of
behaviour of intact rock at different sizes.” PhD, The rocks and concrete under triaxial stress and the effect of
University of New South Wales, Sydney, Australia. pore pressure on the criterion.” Proc., Proceedings 5th Rock
Masoumi, H., Douglas, K. J., and Russell, A. R. (2012). Mechanics Symposium, Pergamon, 563-577.
“Experimental investigation of the size effect of Gosford
Sandstone.” Proc., ANZ 2012 Conference Proceedings, 644-649. Nayak, G. C., and Zienkiewicz, O. C. (1972). “Convenient
Mathey, M., and van der Merwe, J. N. (2015). “The mechanical form of stress invariants for plasticity.” Journal of Structural

296
14 | References

Engineering, 98(ST4), 949-954. Mining Sciences.


Nejati, H. R., Ghazvinian, A., Moosavi, S. A., and Sarfarazi, V. Poulos, H. G., and Davis, E. H. (1974). Elastic solutions for soil
(2014). “On the use of the RMR system for estimation of rock and rock mechanics, John Wiley & Sons.
mass deformation modulus.” Bulletin Engineering Geology and Poulsen, B. A., and Adhikary, D. P. (2013). “A numerical study of
the Environment, 73(2), 531-540. the scale effect in coal strength.” International Journal of Rock
Nemcik, J., Gale, W., and Fabjanczyk, M. W. (2006). “Methods Mechanics and Mining Sciences, 63, 62-71.
of Interpreting Ground Stress Based on Underground Stress.” Poulsen, B. A., Shen, B., Williams, D. J., Huddlestone-Holmes,
Proc., Coal 2006: Coal Operator’s Conference, University of C., Erarslan, N., and Qin, J. (2014). “Strength reduction on
Wollongong & AusIMM, 104-112. saturation of coal and coal measures rocks with implications
for coal pillar strength.” International Journal of Rock Mechanics
Oliveira, D. A. F. (2014). “An alternative view on geotechnical and Mining Sciences, 71, 41-52.
parameters for tunnel design in Sydney.” Australian Priest, S. D. (2005). “Determination of Shear Strength and Three-
Geomechanics, 49(3). dimensional Yield Strength for the Hoek-Brown Criterion.”
Ou, C.-Y. (2006). Deep Excavation: theory and practice, Taylor & Rock Mechanics and Rock Engineering, 38(4), 299-327.
Francis, Leiden, The Netherlands. Priest, S. D., and Brown, E. T. (1983). “Probabilistic stability
analysis of variable rock slopes.” Transactions of the Institution
Palmström, A. (2005). “Measurements of and correlations of Mining and Metallurgy, Section A: Mining industry, 92, A1-
between block size and Rock Quality Designation (RQD).” A12.
Tunnelling and Underground Space Technology, 20, 362-377. Priest, S. D., and Hudson, J. A. (1976). “Discontinuity spacings
Palmström, A. (1995). “RMi – A rock mass characterisation in rock.” International Journal of Rock Mechanics, Mining
system for rock engineering purposes.” PhD, University of Sciences & Geomechanics Abstracts, 13, 135-148.
Oslo, Norway. PSM (2005). “Fatality in Cross City Tunnel.” Unpublished report,
Palmström, A., and Broch, E. (2006). “Use and misuse of rock Carroll & O’Dea.
mass classification systems with particular reference to the Pye, W. (2009). “Rock mass modelling.” Masters of Engineering
Q-system.” Tunnelling and Underground Space Technology, 21, Science, University of New South Wales, Sydney.
575-593.
Palmström, A., and Singh, R. (2001). “The deformation modulus Rafiai, H. (2011). “New empirical polyaxial criterion for rock
of rock masses.” Tunnelling and Underground Space Technology, strength.” International Journal of Rock Mechanics and Mining
16(3), 115-131. Sciences, 48(6), 922-931.
Pan, X. D., and Hudson, J. A. (1988). “A simplified three Rafiai, H., Jafari, A., and Mahmoudi, A. (2013). “Application
dimensional Hoek-Brown yield criterion.” Proc., Rock of ANN-based failure criteria to rocks under polyaxial stress
Mechanics and Power Plants, Balkema, Rotterdam, 95-103. conditions.” International Journal of Rock Mechanics and
Paterson, M. S. (1982). “Biographical memoirs: John Conrad Mining Sciences, 59, 42-49.
Jaeger 1907-1979.” Historical Records of Australian Science, Read, R. S. (1994). “Interpreting excavation-induced
5(3), 1-12. displacements around a tunnel in highly stressed granite.”
Paterson, M. S., and Wong, T. F. (2005). Experimental rock PhD, University of Manitoba, Winnipeg, Manitoba, Canada.
deformation; the brittle field., Springer-Verlag Berlin. Read, S. A. L., and Richards, L. R. (2015). “Guidelines for use of
Patton, F. D. (1966). “Multiple modes of shear failure in rock.” tensile data in the calculation of the Hoek-Brown constant,
Proc., Proceedings 1st Congress International Society of Rock mi.” Proc., 13th International Symposium on Rock Mechanics -
Mechanics, 509-513. ISRM Congress 2015.
Pells, P. J. N. (1993). “Uniaxial strength testing”,3, Comprehensive Reeves, M. J. (1985). “Rock surface roughness and frictional
Rock Engineering, J. A. Hudson, ed., Pergamon, Oxford, 67- strength.” International Journal of Rock Mechanics, Mining
85. Sciences & Geomechanics Abstracts, 22(6), 429-442.
Pells, P. J. N. (2004). “Substance and mass properties for Robertson, A. M. (1970). “The interpretation of geological factors
the design of engineering structures in the Hawkesbury for use in slope theory.” Proc., Planning Open Pit Mines, South
Sandstone.” Australian Geomechanics, 39(3), 1-21. African Institute of Mining and Metallurgy, 55-71.
Pells, P. J. N. (2008). “What happened to the mechanics in rock Robson, R. (1978). “Aspects of testing and sedimentary rock
mechanics and the geology in engineering geology?” Proc., 6th classification for engineering purposes in the Sydney Basin.”
International Symposium on Ground Support in Mining and MSc University of Sydney.
Civil Engineering Construction. Rosso, R. S. (1976). “A comparison of joint stiffness
Pells, P. J. N., Best, R. J., and Poulos, H. G. (1994). “Design of measurements in direct shear, triaxial compression and insitu.”
roof support of the Sydney Opera House parking station.” International Journal of Rock Mechanics, Mining Sciences &
Tunnelling and Underground Space Technology, 9(2), 201-207. Geomechanics Abstracts, 13, 167-172.
Pells, P. J. N., Douglas, D. J., Rodway, B., Thorne, C., and Russo, G. (2009). “A new rational method for calculating the
McMahon, B. K. (1978). “Design loadings for foundations GSI.” Tunnelling and Underground Space Technology, 24, 103-
on shale and sandstone in the Sydney region.” Australian 111.
Geomechanics Journal, G8, 31-39.
Pells, P. J. N., McMahon, B. K., and Redman, P. G. (1980). Saenger, E. H., Kruger, O. S., and Shapiro, S. A. (2004).
“Interpretation of field stresses and deformation moduli “Effective elastic properties of randomly fractured soils: 3D
from extensometer measurements in rock tunnels.” Proc., 4th numerical experiments.” Geophysical Prospecting, 52, 183-195.
Australian Tunnelling Conference, 171-176. Sakurai, S., and Serata, S. (2013). “Mechanical characteristics of
Pells, P. J. N., Mostyn, G. R., and Walker, B. F. (1998). rock salt (3D triaxial).” Proc., True Triaxial Testing of Rocks,
“Foundations on sandstone and shale in the Sydney region.” CRC Press, 181-192.
Australian Geomechanics, 33(2), 17-29. Salamon, M. D. G., Canbulat, I., and Ryder, J. A. (2006). “Seam-
Perras, M. A., and Diederichs, M. (2014). “A review of the specific pillar strength formulae for South African collieries.”
tensile strength of rock: concepts and testing.” Geotechnical & Proc., 41st US Symposium on Rock Mechanics, ARMA/ USRMS.
Geological Engineering, 32(1), 525-546. Salamon, M. D. G., and Munro, A. H. (1967). “A study of the
Potyondy, D., and Cundall, P. A. (2004). “A bonded-particle strength of coal pillars.” Journal of the South African Institute of
model for rock.” International Journal of Rock Mechanics and Mining and Metallurgy (September).

297
ESTIMATING ROCK MASS PROPERTIES

Sari, M. (2012). “An improved method of fitting experimental state mechanics based strength criterion for inherently
data to the Hoek–Brown failure criterion.” Engineering anisotropic rocks.” Proc., 13th International Symposium on Rock
Geology, 127, 27-35. Mechanics - ISRM Congress 2015.
Saroglou, H., and Tsiambaos, G. (2008). “A modified Hoek– Singh, M., and Singh, B. (2012). “Modified Mohr–Coulomb
Brown failure criterion for anisotropic intact rock.” criterion for non-linear triaxial and polyaxial strength of
International Journal of Rock Mechanics and Mining Sciences, jointed rocks.” International Journal of Rock Mechanics and
45(2), 223-234. Mining Sciences, 51, 43-52.
SCA (2012). “Guidelines for development adjacent to the Upper Sloan, S. W., and Booker, J. R. (1986). “Removal of
Canal and Warragamba Pipelines.” Sydney Catchment singularities in Tresca and Mohr-Coulomb yield functions.”
Authority, 24. Communications in Applied Numerical Methods, 2, 171-179.
Schmermann, J. H., and Osterberg, J. H. (1960). “An Soder, P., and Krauland, N. (1989). “Determination of pillar
experimental study of the development of cohesion and strength by full scale pillar tests in the Laisvall Mine.”
friction with axial strain in saturated cohesive soils.” Proc., Proc., Proceedings of the 11th Plenary Scientific Session of the
Research Conference on Shear Strength of Cohesive Soils, International Bureau of Strata Mechanics, Strata Control in Deep
Univerisity of Colorado, 643-694. Mines, Balkema, 39-59.
Seedsman, R. W. (2012). “The strength of the pillar-floor Sonmez, H., and Ulusay, R. (1999). “Modifications to the
system.” Proc., 12th Coal Operators’ Conference, University geological strength index (GSI) and their applicability to
of Wollongong and the Australian Institute of Mining & stability of slopes.” International Journal of Rock Mechanics and
Metallurgy, 23-30. Mining Sciences, 36, 743-760.
Seedsman, R. W., Jalalifar, H., and Aziz, N. (2005). “Chain Sonmez, H., and Ulusay, R. (2002). “A discussion on the Hoek–
Pillar Design - Can We?” Proc., Coal Operators’ Conference, Brown failure criterion and suggested modifications to the
University of Wollongong & the Australasian Institute of criterion verified by slope stability case studies.” Yerbilimleri,
Mining and Metallurgy, 59-62. 26, 77-99.
Seidel, J. P., and Haberfield, C. M. (2002). “A theoretical model Stacey, T. R. (1981). “A simple extension strain criterion for
for rock joints subjected to constant normal stiffness direct fracture of brittle rock.” International Journal of Rock Mechanics,
shear.” International Journal of Rock Mechanics and Mining Mining Sciences & Geomechanics Abstracts, 18, 469-474.
Sciences, 39, 539-553. Stacey, T. R. (2000). “Reservations regarding the use of rock
Serafim, J. L., and Pereira, J. P. (1983). “Considerations on the mass classifications in rock engineering.” Proc., Proceedings
geomechanical classification of Bieniawski.” Proc., International Bergmekanikdag 2000, SveBeFo, Swedish Rock Engineering
Symposium on Engineering Geology and Underground Research and National Group ISRM, 1-18.
Construction, International Association of Engineering Stacey, T. R., and Page, C. H. (1986). Practical handbook
Geology, 33-42. for underground rock mechanics, Trans Tech Publications,
Serrano, A., Estaire, J., and Olalla, C. (2007). “Extension of the Germany.
Hoek-Brown failure criterion to three dimensions.” Proc., 11th Stephens, R. E., and Banks, D. C. (1989). “Moduli of
Congress of the ISRM, Taylor & Francis, Leiden, 289-292. deformation studies of the foundation and abutments of the
Sharp, J. C., Bandis, S. C., Schinas, C. A., MacKean, R. N., Portugues Dam - Puerto Rico.” Proc., Rock mechanics as a guide
and Watson, S. P. (2008). “Engineering evaluation of design for efficient utilization of natural resources, Balkema, 31 - 39.
concepts for a large span urban underground cavern in weak Swarbrick, G. ‘Devines Tunnel Mapping’. PSM, 2013.
rock based on design analysis.” Proc., Continuum and Distinct SWC-Transwater (1996). “Cataract Tunnel - geological inspection
Element Numerical Modelling in Geo-Engineering, Itasca. - 10/4/1996.” 12.
Shen, B., Nash, T. R., Bertuzzi, R., and Clarke, S. J. (2015).
“Use of monitoring data during construction to refine Takahashi, M., and Kiode, H. (1989). “Effect of intemediate
cavern design.” Proc., 9th International Symposium on principal stress on strength and deformation behaviour of
Field Measurements in Geomechanics, Australian Centre for sedimentary rocks at the depth shallower than 2000m.” Proc.,
Geomechanics. Rock at Great Depth, Balkema, 19-26.
Sheorey, P. R. (1997). Empirical rock failure criteria, Balkema, Tarasov, B., and Potvin, Y. (2013). “Universal criteria for
Rotterdam. rock brittleness estimation under triaxial compression.”
Sheorey, P. R., Biswas, A. K., and Choubey, V. D. (1989). “An International Journal of Rock Mechanics and Mining Sciences,
empirical failure criterion for rocks and jointed rock masses.” 59, 57-69.
Engineering Geology, 26, 141-159. Taylor, J. A., and Fowell, R. J. (2003). “Estimating the in-situ
Sheorey, P. R., Das, M. N., Barat, D., Prasad, R. K., and Singh, B. mass strength and time to failure of coal pillarsin bord-and-
(1987). “Coal pillar strength estimation from failed and stable pillar workings.” Proc., ISRM 2003 - Technology roadmap
cases.” International Journal of Rock Mechanics, Mining Sciences for rock mechanics, South African Institute of Mining and
and Geomechanical Abstracts, 24(6), 347-355. Metallurgy, 1203-1208.
Sheorey, P. R., and Singh, B. (1982). “Application of a rock mass Terzaghi, K. (1946). “Load on tunnel supports”,4, Rock tunneling
classification to mining stability problems - some case studies.” with steel supports, R. V. Proctor, and T. L. White, eds., The
Proc., Proceedings of the First International Conference on Commercial Shearing and Stamping Company, Youngstown,
Stability in Underground Mining, AIME, 383-395. Ohio, 47-86.
Singh, J., Ramamurthy, T., and Venkatappa Rao, G. (1989). Thorpe, R., Watkins, D. J., Ralph, W. E., Hsu, R., and Flexser, S.
“Strength anisotropies in rocks.” Indian Geotechnical Journal, (1980). “Strength and permeability tests on ultra-large Stripa
19(2), 147-166. granite core.” Lawrence Berkeley Laboratory, University of
Singh, M., Raj, A., and Singh, B. (2011). “Modified Mohr– California, Berkeley, CA, 233.
Coulomb criterion for non-linear triaxial and polyaxial Tinsley, C. R. (1982). “Supply and Demand for Coal”,15,
strength of intact rocks.” International Journal of Rock Economics of Pacific Rim coal, 113-122.
Mechanics and Mining Sciences, 48(4), 546-555. Townsend, J. M., Jennings, W. C., Haycocks, C., Neall, G.
Singh, M., Samadhiya, N. K., Kumar, A., Kumar, V., and M., and Johnson, L. P. (1977). “A relationship between the
Singh, B. (2015). “A nonlinear criterion for triaxial strength ultimate compressive strength of cubes and cylinders for coal
of inherently anisotropic rocks.” Rock Mechanics and Rock specimens.” Proc., 18th US Symposium on Rock Mechanics,
Engineering, 48(4), 1387-1405. Colorado School of Mines Press, 4A6-1 - 4A6-6.
Singh, M., Samadhiya, N. K., and Singh, B. (2015). “Critical Tziallas, G. P., Tsiambaos, G., and Saroglou, H. (2009).

298
14 | References

“Determination of rock strength and deformability of Yoshinaka, R., and Yamabe, T. (1980). “Strength criterion of
intact rocks.” Electronic Journal of Geotechnical Engineering, rocks.” Soils and Foundations, 20(4), 113-126.
14(Bundle G), 1-12. You, M. (2009). “True-triaxial strength criteria for rock.”
Unal, E. (1983). “Development of design guidelines and roof- International Journal of Rock Mechanics and Mining Sciences,
control standards for coal-mine roofs.”PhD, Pennsylvania State 46(1), 115-127.
University. You, M. (2011). “Comparison of the accuracy of some
conventional triaxial strength criteria for intact rock.”
van der Merwe, J. N. (1993). “Revised strength factor for coal in International Journal of Rock Mechanics and Mining Sciences,
the Vaal basin.” Journal of the South African Institute of Mining 48(5), 852-863.
and Metallurgy, 93(3), 71-77. You, M. (2012). “Comparison of two true-triaxial strength
van der Merwe, J. N. (2003). “New pillar strength formula for criteria.” International Journal of Rock Mechanics and Mining
South African coal.” The Journal of The Southern African Sciences, 54, 114-124.
lnstitute of Mining and Metallurgy, June 2003, 281-292. You, M. (2015). “Strength criterion for rocks under compressive-
van der Merwe, J. N., and Mathey, M. (2013). “Update of coal tensile stresses and its application.” Journal of Rock Mechanics
pillar database for South African coal mining.” The Journal of and Geotechnical Engineering.
The Southern African Institute of Mining and Metallurgy, 113, Yu, H. (2006). “Plasticity and geotechnics”, Advances in mechanics
825-840. and mathematics, D. Y. Gao, and R. W. Ogden, eds., Springer,
van Heerden, W. L. (1975). “In situ complete stress-strain New York.
characteristics of large coal specimens.” Journal of the South
African Institute of Mining and Metallurgy(March 1975), 207- Zhang, L. (2010). “Method for estimating the deformability
217. of heavily jointed rock masses.” Journal of Geotechnical and
Vernik, L., and Kachanov, M. (2012). “On some controversial Geoenvironmental Engineering, 136(9), 1242-1250.
issues in rock physics.” The Leading Edge, 636-642. Zhang, L., and Einstein, H. H. (2004). “Using RQD to estimate
Vesic, A. S. (1961). “Beams on elastic subgrade and the Winkler the deformation modulus of rock masses.” International
hypothesis.” Proc., 5th International Conference on Soil Journal of Rock Mechanics and Mining Sciences, 41, 337-341.
Mechanics and Foundation Engineering, 845-850. Zhang, L., and Einstein, H. H. (2009). “The Planar Shape of
Vibert, C., and Ianos, S. (2015). “Moving towards a reliable Rock Joints.” Rock Mechanics and Rock Engineering, 43(1),
assessment of deformability of rock masses: examples from 55-68.
large dams foundations.” Proc., 13th International Symposium Zhang, L., and Zhu, H. (2007). “Three-dimensional Hoek-Brown
on Rock Mechanics - ISRM Congress 2015. strength criterion for rocks.” Journal of Geotechnical and
Geoenvironmental Engineering, 133(9), 1128-1135.
Wagner, H. (1974). “Determination of the complete load- Zhang, Q., Zhu, H., and Zhang, L. (2013). “Modification of a
deformation characteristics of coal pillars.” Proc., Advances in generalised three-dimensional Hoek-Brown strength criterion.”
rock mechanics, Proceedings of the 3rd Congress International International Journal of Rock Mechanics and Mining Sciences,
Society for Rock Mechanics, 1076-1081. 59, 80-96.
Wallis, S. (2000). “Sydney’s Northside Storage Tunnel Zhou, S. (1994). “A program to model the initial shape and extent
alliance holding up under pressure.” Tunnels & Tunnelling of borehole breakout.” Computers and Geosciences, 20(7/8),
International, 37-40. 1143-1160.
Wallis, S. (2007). “Australia examines toll road collapses.” Tunnels Zimmerman, R. W. (1985). “The effect of microcracks on the
& Tunnelling International, 26-29. elastic moduli of brittle materials.” Journal of Materials Science
Walsh, J. B. (1965). “The effect of cracks in rocks on Poisson’s Letters, 4, 1457-1460.
Ratio.” Journal of Geophysical Research, 70(20), 5249-5257. Zimmerman, R. W. (1991). Compressibility of sandstones, Elsevier
Walsh, J. B. (1965). “The effect of cracks on the compressibility of Science Publishers.
rock.” Journal of Geophysical Research, 70(2), 381-389. Zipf, R. K. (1999). “Using a post-failure stability criterion in pillar
Walsh, J. B., Brace, W. F., and W, E. A. (1965). “Effect of porosity design.” Proc., Proceedings of the Second International Workshop
on compressibility of glass.” Journal of the American Ceramic on coal pillar mechanics and design, 181-192.
Society, 48(12), 605-608. Zuo, J.-p., Li, H.-t., Xie, H.-p., Ju, Y., and Peng, S.-p. (2008). “A
Wiebols, G. A., and Cook, N. G. W. (1968). “An energy nonlinear strength criterion for rock-like materials based on
criterion for the strength of rock in polyaxial compression.” fracture mechanics.” International Journal of Rock Mechanics
International Journal of Rock Mechanics and Mining Sciences, 5, and Mining Sciences, 45, 594-599.
529-549. Zuo, J., Liu, H., and Li, H. (2015). “A theoretical derivation of
Wilson, J. W. (1971). “The design and support of underground the Hoek-Brown failure criterion for rock materials.” Journal of
excavations in deep-level hard rock mines.” PhD, University of
the Witwatersrand, Johannesburg, South Africa.
Won, G. W. (1985). “Engineering properties of Wianamatta
Group rocks from laboratory and in-situ tests”, Engineering
Geology of the Sydney Region, P. J. N. Pells, ed., AA Balkema,
Rotterdam, 143-155.

Yavuz, H., and Fowell, R. J. (2001). “FDM prediction of a yield


pillar performance in conjunction with a field trial.” Proc.,
20th International Conference on Ground Control in Mining,
University of West Virginia, 78-85.
York, G., Canbulat, I., and Jack, B. W. (2000). “Coal pillar design
procedures.” S. i. M. R. A. Committee, ed., CSIR Mining
Technology, 216.
Yoshinaka, R., Osada, M., Park, H., Sasaki, T., and Sasaki,
K. (2008). “Practical determination of mechanical design
parameters of intact rock considering scale effect.” Engineering
Geology, 96(3-4), 173-186.

299
ESTIMATING ROCK MASS PROPERTIES

Appendices

ROCK MASS PROPERTIES FOR TUNNELING

CONTENTS
A Development of Barton shear strength criterion

B Development of Hoek-Brown strength criterion

C Development of Lade criterion

D Development of Christensen criterion

300
14 | References

A: Development of Barton Criterion

CRITERION At lower stresses, the normal stress could be


calculated from considerations of Mohr circles. That
In the late 1960s and early 1970s, Barton researched
is, the theoretical relationships between σn, τ and the
the strength and stiffness of rock joints and developed
principal stresses are:
an empirical relationship for joint strength e.g. Barton
(1973). Barton fitted that empirical relationship to the
published triaxial strength data of three intact rocks –
Solenhofen limestone, Spruce Pine dunite and Westerly
granite – and proposed the following relationship to
estimate the shear strength of the mobilised failure
plane of the triaxial samples (Barton 1976).

EQUATION A1

As an aside, Barton refers to the differential stress at


failure (σ1-σ3) as the intact rock’s “confined strength”.
Equation A1 is in the form of the Mohr-Coulomb
criterion without the cohesive intercept and with a
variable friction angle. That is, it is in the form τ =
σn tanψ. The equation is invalid for stress conditions
where the log10 term, (σ1 – σ3) / σn, is negative. It also
becomes invalid when angle ψ approaches 90°. Re-
phrasing, Equation A1 is valid therefore when:

and

In developing the empirical relationship in Equation


A1, Barton (1976) assumed σn = 2σ3 was realistic for
‘medium stress levels’. Further, Barton proposed that
the maximum shear strength is reached at the ‘critical
state’, which corresponds to σn = 2σ3, τ = σ3 and ‘a
satisfactory approximation’ of σ1 = 3σ3.

301
ESTIMATING ROCK MASS PROPERTIES

The angle β is that which the failure plane makes Equation A1 can be fitted to experimental intact
with the major principal stress, σ1 – see margin sketch rock data to determine the shear strength, τ, given
adapted from Brady and Brown (2005). Barton (1976) the principal stresses, σ1 and σ3. However, unlike the
notes that a value of β = 30° is a reasonable mean, Mohr-Coulomb criterion, the ‘friction angle’ (ψ) and its
with experimental values typically between 20 to 39°. ‘conjugate angle’ (β) are variable; themselves functions
Though as the shear strength envelope is curved as of σ1 and σ3. Hence, Equation A1 and Equation A2
shown in Figure A1, β increases with confinement, do not predict a unique maximum principal stress at
σ3, and hence is really a variable. Barton (1976) also failure for a particular confinement.
assumed the shear strength envelope is a straight line In order to fit Equation A1 to laboratory data of
in the low compressive stress environment and in intact rock, Barton (1976) used the strength criterion
the tensile stress environment, i.e. between the UCS of Bieniawski (1974) given in Equation A3 and adopted
and UTS. k = 0.75 and A between 3 and 5.
The general form of the Barton shear strength
criterion, Equation A2, has been proposed for the
strength of intact rock, jointed rock, rock fill and
infilled defects as shown in Figure A1, e.g. Barton
EQUATION A3
(1999); (2014).

Barton (1976) then developed an empirical method


to derive the shear (τ) and normal (σn) stresses from the
Bieniawski criterion and showed that Equation A2 fits.
EQUATION A2
The empirical method to derive τ and σn from the
Bieniawski criterion (Equation A3) is detailed in the
Where:
following steps and summarised in Equation A4.
X represents roughness
Y represents the effective compressive strength • Equation A3 is used to calculate (σ1 - σ3) for
various σ3. The normal stress is assumed equal to
Z represents the basic and residual friction. σn = 2σ3 and hence the shear stress at failure (τ) is
estimated with Equation A1 (Barton 1976).
For intact rock, X = 50, Y = and Z = ϕc = 26.6°, • For low stresses, the shear strength envelope
i.e. Equation A1. is assumed to be linear from the ‘cohesive
intercept’, (Barton 1976).

FIGURE A1 Shear strength relationships for intact rock through to jointed rock, rock fill and infilled defects as proposed by
Barton, e.g. (Barton 1999; 2014)

302
Appendices – Rock Mass Properties for Tunnelling • Appendix A: Development of Barton Criterion

• Further, the principal stresses can be inferred by re-


arranging the Mohr circle relationship as follows

EQUATION A4

The author suggests the following alternative. It is a


multi-step, trial and error approach to initially fit the
criterion to experimental data. The fitted criterion can
then be used to predict the strength of intact rock.
• Using the laboratory data (σ3, σ1), values of normal
stress (σn) are found assuming the Mohr circle
relationships given previously and β = 30°. This is
consistent with Barton (1976).
• The ‘friction angle’ (ψ) can be calculated by noting
from Equation A1 that:
hence

EQUATION A5

The relationship between ψ and σn is plotted and


a smooth, continuous line of best fit is found. This
allows values of ψ to be calculated for any σn and by
extension any σ3. The author has found that the line
of best fit generally follows a power equation of the
form: ψ = Cσ3-D.
• Fit Equation A3 to the laboratory data (σ3, σ1) by
selecting the parameter A to minimise the error
between σ3 predicted by and that measured, i.e.

minimise Σ(σ1predicted - σ1measured)2.
• The differential stress (σ1predicted - σ3) is calculated.
• The normal stress (σn) is calculated by re-arranging
Equation A5 as:

EQUATION A6

• The shear stress (τ) is calculated using Equation A1


• For interest, the Mohr circle relationship can now
be used to find the values of β that apply along the
curved criterion.

303
ESTIMATING ROCK MASS PROPERTIES

B: Development of Hoek-Brown
Criterion

CRITERION Squaring both sides and simplifying


It is worth re-visiting the original work of Hoek
(1965) and Hoek and Bieniawski (1965) to show the
development of the Hoek-Brown criterion from the
Griffith crack extension theory. The Griffith’s crack
theory was written in terms of σ1, σ3 and σc as follows.

EQUATION B1
which by making the prefixes to σc the variables m and
Re-arranging Equation B1 to suit a quadratic solution: s (notionally)

became the original Hoek-Brown criterion (Hoek and


Brown 1980):

which gives solutions of:

EQUATION B2

The parameters m and s were defined initially by


Substituting σ1/8= – σ1, which is a fixed relationship Hoek and Brown (1980) for intact rock as m = mi
for Griffith’s criterion ≈ σc / |σt| and s = 1.0. The generalised Hoek-Brown
criterion introduced by Hoek (1994) and then Hoek et
al. (2002) made the parameter a variable between 0.5
and 0.65 (Equation B3).

EQUATION B3

304
Appendices – Rock Mass Properties for Tunnelling • Appendix B: Development of Hoek-Brown Criterion

In considering the development of micro-cracks in However, Zhang and Zhu (2007) noted that
intact rock, Zuo et al. (2008) and later extended by Equation B4 does not simplify to the Hoek-Brown
Zuo et al. (2015) derive a strength criterion which can criterion in conventional triaxial tests, i.e. when σ2=
be written in the same format as the original Hoek- σ3. Hence, they instead proposed that the mean
Brown criterion: normal stress (or octahedral stress σoct) be restricted to
consider only σ1 and σ3.
By not including σ2 into the calculation of the mean
normal stress, Zhang and Zhu (2007) followed the
observation initially made by Mogi (1967) that the
initial failure path in tested samples occurred in the
where σ2 direction and hence the normal stress acting on
that failure path was a function of only σ1 and σ3. The
 is defined as the friction coefficient of the
μ octahedral stress σoct could effectively be replaced by
micro-crack the mean normal stress σm,2 =(σ1+σ3)/2.

 is the ‘mixed-mode’ fracture coefficient and


κ
ranges from √3/2 to 1 to
depending if a stress, energy or strain criteria is EQUATION B5
assumed.
The Zhang and Zhu (2007) modification does simplify
Zuo et al. (2008) state that the parameter m can to the Hoek-Brown equation in the triaxial (σ2 = σ3)
characterise the rock. and biaxial (σ1 = σ2) states. In triaxial state where σ2 =
σ3, Equation B5 simplifies to
Assuming the typical value of Poisson’s Ratio for
intact rock is v = 0.25, κ is implied to vary between
0.866 – 1.240 and hence β between 0.758 – 1.085.
Zuo et al. (2015) quoting Paterson and Wong
(2005) state μ varies between 0.2 – 0.8. This range for
μ implies a wide range for the friction angle of clean
micro-cracks in rock of between 11 – 39°.
For these ranges of β and μ, the ratio μ/β varies
between 0.184 – 1.056. That is,
This in turn suggests that the Hoek-Brown
parameter, m≈(0.2 to 1.0)× .

3D EXTENSIONS
Several extensions to the Hoek-Brown criterion to
explicitly consider three dimensional stresses have
been proposed. One of the earliest was that of Pan and
EQUATION B6
Hudson (1988) who following the work of Nayak and
Zienkiewicz (1972) wrote the criterion in terms of the
stress invariants and derived the following equation. Equation B6 is the 2D Hoek-Brown criterion shown
in Equation B2.
Zhang et al. (2013) note that the 3D failure surface
generated by Equation B5 is not smooth and convex
in all the stress states. To overcome this limitation,
EQUATION B4 they introduce a Lode angle function that is iteratively
solved. Similar approaches have been used to ensure
where smooth and convex failure envelopes of other 3D
criteria, e.g. Sloan and Booker (1986).
A similar 3D development of the Hoek-Brown
criterion was followed by Jiang et al. (2011) who made
the following substitutions, noting that σm =I1.

305
ESTIMATING ROCK MASS PROPERTIES

Substituting these into the Hoek-Brown criterion


given in Equation B2 we get:

Collecting like terms and simplifying

Noting that

Squaring both sides

Noting that

EQUATION B7

Equation B7 is the Hoek-Brown criterion in the


format given in Jiang et al. (2011).

306
Appendices – Rock Mass Properties for Tunnelling • Appendix C: Development of Lade Criterion

C: Development of Lade Criterion

This appendix develops equations to calculate the Lade Lade’s criterion can be written as (Lade 1993):
parameters a, m, η1, from the standard laboratory tests
– UCS and the Brazilian Tensile Strength (BTS).

CRITERION EQUATION C8

Lade tested the strength of soil under uniaxial, biaxial,


where
triaxial and torsional stresses in a laboratory (Lade and
Duncan 1973) and developed an elasto-plastic stress-
strain theory for cohesionless soils (Lade 1977; Lade
and Duncan 1975). He later extended this to concrete
and rock (Kim and Lade 1984; Lade 1982; Lade 1984).
His observations led him to conclude that a • The value of I13 / I3 is 27 at the hydrostatic axis, i.e.
three-dimensional failure surface has four main σ1 = σ2 = σ3
characteristics as shown in Figure C1.
• The term apa reflects the tensile strength of the
• It is a smooth, triangular, monotonically curved material. It is slightly greater than the UTS (Figure
failure surface in the octahedral plane C1). Lade (1993) uses values of a that are 0.1 to 2%
• The opening angle of the failure surface is described higher than σt /pa.
by the material’s friction angle, as in the Mohr- • The apex angle of the failure surface increases with
Coulomb criterion the value of η1. The curvature of the failure surface
• The failure surface is curved in planes containing increases with the value of m. See Figure C2.
the hydrostatic axis • The criterion assumes the material behaves
• It can have a tensile strength. isotropically.

FIGURE C1 3D failure surface viewed (a) from the side to show curved envelope in triaxial plane and biaxial planes and (b)
from the end to show smoothly rounded, triangular shape in octahedral plane (Lade 1993)

307
ESTIMATING ROCK MASS PROPERTIES

FIGURE C1 Translation of principal stress space along the hydrostatic axis to include the effect of tensile strength in the
failure criterion (Lade 1993)

FIGURE C2 Influence of parameters m, η1 and a (Lade 1993)

308
Appendices – Rock Mass Properties for Tunnelling • Appendix C: Development of Lade Criterion

Estimates of the three parameters a, η1 and m are • Sedimentary


required to use the criterion and can be obtained  a = 3.0 to 230; m = 0.502 to 1.93;
following the two straightforward steps (Lade 1993): η1 = 4.5 x103 to 1.5x109
• An estimate of a is made first then apa is added to the • Coal
normal stresses before substitution into Equation B1.  a = 3.6 to 32; m = 0.520 to 1.28;
• 3 e parameters η1 and m can be obtained by plotting
Th
 η1 = 3.9 x103 to 4.7x105
I1 / I3 – 27 versus pa/ I1 on a log-log diagram and
locating the best fitting straight line.
The ranges of the parameters are very large. The
• The intercept of this straight line with pa/ I1 =1 is the range for η1 spans seven orders of magnitude from 540
value of η1 and its slope is m.
to 150,000,000. The value of a varies by two orders of
However, an infinite number of combinations of a, magnitude; from 3 to over 400. The values for m are
η1 and m can result in the same value of the unconfined more constrained; varying between 0.3 and 2.5.
compressive strength, σc. Lade (1993) studied 87 sets
of laboratory data for igneous (19 sets), metamorphic Lade (1993) concluded that typical values for rock are:
(12), sedimentary (33) and coal (23) rocks and found • m = 0.7 to 1.9 and η1 = 104 to 108 (Figure C3)
the following range of values of the parameters. • a = 20 to 300 except for coal which has a = 3 to 30
(Figure C4)
• I gneous
 a = 2.9 to 410; m = 0.296 to 2.50;
η1 = 5.4 x102 to 3.6x108
• Metamorphic
 a= 9.1 to 303; m = 0.667 to 1.93;
η1 = 4.3 x104 to 2.8x108

FIGURE C3 Variation and range of η1 and m for various frictional materials (Lade 1993)

309
ESTIMATING ROCK MASS PROPERTIES

FIGURE C4 Variation and range of η1 and a relative to σt for various frictional materials (Lade 1993)

Critically looking at the intact rock test data published example. The typical scatter of laboratory test results
by Lade, for example Lade (1993), it appears that does not appear in these figures. Hence, the published
generally one or only a few test values of (σt , σc) and test data may give a false sense that the criterion is very
(σ1, σ3) pairs were used to estimate the parameters a, η1 well-conditioned.
and m, see the following Figure C5 to Figure C11 for

FIGURE C5 Results of triaxial, biaxial and cubical triaxial compression tests on sandstone projected onto an octahedral plane
(Lade 1993)

310
Appendices – Rock Mass Properties for Tunnelling • Appendix C: Development of Lade Criterion

FIGURE C6 Results of biaxial tests on Indiana limestone showing shape of failure surface in biaxial plane (Lade 1993)

FIGURE C7 Results of tests on Mizuho trachyte in (a) triaxial plane and (b) octahedral plane (Lade 1993)

311
ESTIMATING ROCK MASS PROPERTIES

FIGURE C8 Results of tests on dolomite in (a) triaxial plane and (b) octahedral plane (Lade 1993)

FIGURE C9 Results of tests on Wombeyan marble in (a) triaxial plane and (b) octahedral plane ((Lade 1993)

312
Appendices – Rock Mass Properties for Tunnelling • Appendix C: Development of Lade Criterion

FIGURE C10 Results of tests on sandstone in (a) triaxial plane, (b) biaxial plane and (c) octahedral plane (Lade 1993)

FIGURE C11 Results of tests on Darney sandstone in (a) biaxial plane and (b) octahedral plane (Lade 1993)

313
ESTIMATING ROCK MASS PROPERTIES

PARAMETER DEFINITION It can be seen from Equation C9, which is in the


standard form of a straight line, y=mx+b, that the
According to Lade (1984) for a given set of strength
parameters η1 and m can be obtained by plotting I13 /
data, an estimate of the tensile strength parameter a
I3 – 27 versus pa / I1 on a log-log diagram and locating
is made first such that the term apa is slightly greater
the best fitting straight line. The intercept of this line
than the uniaxial tensile strength, σt. Lade (1984);
with pa / I1 =1 is the value of η1 and its slope is m (see
Lade (1993) suggests that apa is within 2% of σt
Figure C13). A line connecting a pair of test results
(Figure C12).
can therefore theoretically be used to define the Lade
The parameter a can be considered to be a function parameters, η1 and m.
of σt. : a=f( σt )
The parameters can also be seen in Figure C13 to be
The value of apa is then added to the normal stresses relatively ill-conditioned. An insignificant change in a
before substitution into Equation B1 and estimates for from 23.30 to 23.15 as shown in the straight-line fits
η1 and m found. Re-arranging Equation B1 and taking to the data needs a 12-fold increase in η1, from 33,821
logs on both sides gives: to 419,812.

EQUATION C9

FIGURE C12 Magnitude of the parameter a relative to σt plotted against η1 from Lade (1984). The red line shows the
majority of the data indicates apa is within 2% of σt .

314
Appendices – Rock Mass Properties for Tunnelling • Appendix C: Development of Lade Criterion

As the slope on a log-log plot gives m and its intercept


gives η1, Lade (2013) suggested that the parameters
are related by Equation C10 and could be estimated
from the results of one laboratory test. He found the
variables A and B given in Equation C10 for the main
groups of rocks to be as listed in Table C1.

EQUATION C10

FIGURE C13 The parameters m and η1 can be obtained from log-log plots, such as this from Lade (1997) showing results
from laboratory tests.

315
ESTIMATING ROCK MASS PROPERTIES

TABLE C1 Values of A and B from Lade (2013)


Rock Type No. of data sets A B r2

Igneous 18 0.249 55.1 0.9265

Metamorphic 12 0.257 22.7 0.8330

Sedimentary 34 0.210 2.15 0.9011

Organic (coal) 22 0.358 130.9 0.9765

All rocks (no coal) 64 0.229 9.62 0.8752

LABORATORY TESTING
Consider the laboratory tests; UCS and UTS and test carried out for rock. Further, Lade (1993) considers
assume that the UCS and UTS values are very good that it is advantageous to include the tensile strength
estimates of the unconfined compressive and tensile in determining the parameters.
rock strengths. That is, the scatter in test results Let (x,y) be the point that represents the UCS test and
is negligible. (x1,y1) the UTS test on the log [I13 / I3 – 27] versus log
The UCS and UTS are chosen by the author as they [pa / I1] graph. The slope of the straight line connecting
straddle the low confinement environment which is these two points is then the Lade parameter m and the
typically the stress state of most interest in tunnelling. intersection of this line with the axis defined by log
The UCS is also the most common laboratory strength (pa / I1) = 0 gives log η1.

FIGURE C14 Translation of principal stress space along the hydrostatic axis to include the effect of tensile strength in the
failure criterion; after Lade (1982)

316
Appendices – Rock Mass Properties for Tunnelling • Appendix C: Development of Lade Criterion

For the UCS test Substituting the values for the points UCS (x,y) and
UTS (x1,y1) into the equation for a straight line, i.e.
y1=y+m(x1-x), the following rather messy equation
is obtained.

EQUATION C11

Hence the point representing the UCS test on the


log [I13 / I3 – 27] versus log [pa / I1] graph is given by This can be somewhat simplified as:
(x, y):

i.e.

Similarly, for the UTS test and to be consistent with EQUATION C13
the terminology used by Lade (Figure C14) it can be
written as: Therefore it can be seen that the parameter m can be
written as a function of σc and σt:

The intercept between the straight line through


points UCS (x,y) and UTS (x1,y1) and the vertical axis
EQUATION C12 defined by log (pa / I1) = 0, i.e. pa / I1 = 1, is the value of
log η1. Let this point on the log – log graph be given
Hence the point representing the UTS test on the log by (x2,y2).
[I13 / I3 – 27] versus log [pa / I1] graph is given by (x1, y1): Substituting (x2,y2) into the equation for a line,
y2=y1+m(x2-x1) gives:

EQUATION C14

So the parameter η1 can also be written as a function


of σc and σt:

317
ESTIMATING ROCK MASS PROPERTIES

Equation C14 could have been written with any Combining Equation C15 with the values in Table
point that lies on the straight line. However, it was C1, pairs of equations for the five rock groups with
purposefully written with the point UTS (x1,y1) as basically the one unknown, σt , can be obtained. First,
we will now take advantage of the fact that apa is an the equation for B gives j as follow.
estimate of σt.
Equation C14 can be re-arranged to be in the same
form as suggested by Lade (2013), i.e.:

Defining variables A and B as: EQUATION C17

Solving this cubic equation for j, then σt can be solved


using the equation for A as follows.

EQUATION C15

Gives:

EQUATION C16

which is the same equation given in Lade (2013), see


Equation C10.
Using the fact that apa must be slightly greater than
the uniaxial tensile strength as noted by Lade (1984)
and shown in Figure C12, then apa = j|σt| where j is
slightly greater than 1.0.

EQUATION C18

TABLE C2 Values of A and B from Lade (2013) with the corresponding predicted tensile strength
Rock Type B j % that A σt
apa > σt [MPa]
Igneous 55.1 1.130 13 0.249 440
Metamorphic 22.7 1.282 28 0.257 277
Sedimentary 2.15 2.563 156 0.210 875
Organic (coal) 130.9 1.058 6 0.358 29
All rocks (no coal) 9.62 1.558 56 0.229 642

318
Appendices – Rock Mass Properties for Tunnelling • Appendix C: Development of Lade Criterion

The solutions to the cubic Equation C17 in j are These approximations flow through such that the
given in Table C2 for the five values of B and range equation for m (Equation C13) becomes:
from 1.058 to 2.563. This means that for the data
considered by Lade (2013) the inferred values of apa
are approximately 6 to 156% greater than σt . This is
a surprising and unexpected result considering Lade
(1993) suggested apa was within 2% of σt (see Figure
C12).
Substituting these values of j into Equation C18
using the corresponding values of A from Table C1
unfortunately predicts tensile strengths which are
extremely high and in the author’s opinion artificially
so, as shown in Table C2.
The surprising and unexpectedly high values of
apa and σt found by the above analysis questions the
usefulness of estimating the Lade parameters from a
single laboratory test as suggested by Lade (1993). EQUATION C19

SUGGESTED METHOD TO
and the equation for η1 (Equation C14) becomes:
ESTIMATE PARAMETERS
The Lade parameter a is considered to be an estimate
of the material’s uniaxial tensile strength. Its value is
chosen such that the term apa is slightly greater than
σt , typically within 2% of σt for rock e.g. Lade (1984);
Lade (1993) as shown in Figure C12. It is possible
therefore to make the approximations apa σt and apa
≈1.02σt .
The points representing the UCS (x,y) and the UTS
(x1,y1) which were found earlier can therefore be
written as:

UCS

EQUATION C20

1.00E+06 2.00
UCS= 1MPa
UCS= 10MPa
UCS= 100MPa
1.00E+05 1.50
UTS UCS= 250MPa
m

η1 1.00E+04 1.00 m

1.00E+03 0.50

1.00E+02 0.00
0 2 4 6 8 10 12 14 16 18 20
σc/σt

FIGURE C15 relationship between Lade parameters and


The log term for y1 is equal to B (see Equation C15). the ratio σc / σt for typical rocks. The parameter η1 is also
Therefore interestingly, when apa = 1.02σt , B ≈ 393. dependent on UCS. Implicit in this graph is that apa = 1.02σt.

319
ESTIMATING ROCK MASS PROPERTIES

D: Development of Christensen
Criterion

This appendix develops the Christensen failure


criterion in much the same way as Hoek and Brown
developed their original criterion from the Griffith
crack extension theory (Hoek and Brown 1980). The
intention is to see whether the Christensen criterion
can be transformed to resemble the style of the Hoek-
Brown criterion.

CRITERION
In a series of papers Christensen (2007) advanced
a failure criterion by considering the stress state
to characterise the elastic energy of a homogenous
material. In this, he started with fracture mechanics
introduced by Griffith’s crack extension. Christensen
assumed that a homogenous material does not fail under
hydrostatic pressure but could fail under hydrostatic
tension. The criterion was developed principally for
use with man-made isotropic materials. It is initially
written here in the form given by Christensen (2007)
with tension assumed to be positive but in terms of
principal stresses.

EQUATION D1

For brittle materials, such as rock, which are defined


by Christensen as having 0 ≤ σt ≤ 0.5|σc|, the maximum
principal tensile stress is limited to a maximum of σt.
The criterion is valid for materials where the ratio
0 ≤ σt /|σc| ≤1. At the limit of σt =|σc|, that is very
ductile behaviour, the Christensen criterion becomes
the von Mises criterion. The criterion can differentiate
between ductile and brittle failure.
Adopting the standard rock mechanics convention,
which has compression stresses being positive,
Equation B1 is re-written as:

EQUATION D2

320
Appendices – Rock Mass Properties for Tunnelling • Appendix D: Development of Christensen Criterion

Equation B1 can be re-arranged as follows.

Hence, the criterion can be written as:

EQUATION D3

The solution to Equation D3 can be found by first


expanding and then simplifying as follows.

Substituting variables b and c as:

EQUATION D4

321
ESTIMATING ROCK MASS PROPERTIES

The following quadratic equation is obtained.

From which solutions for the major principal stress in


Equation D3 can be found as:

VALIDATION
The solutions to the Christensen criterion for
conventional triaxial tests, σ2 = σ3 and |σt| < 1/2σc, can
be found by substituting for the constants b and c as:

Hence the solutions for σ1 for conventional triaxial


tests are given by:

For a UCS test, σ2 = σ3 = 0, the solutions further


simplify to:

Squaring both sides

322
Appendices – Rock Mass Properties for Tunnelling • Appendix D: Development of Christensen Criterion

Simplifying

which has roots

and the expected result of

HOEK-BROWN STYLE
DEVELOPMENT
Equation D3, which is repeated here,

is of a similar form to the Griffith crack extension


theory as extended by Murrell (1963):

The similarity of form between the equations is not


surprising given Christensen used fracture mechanics
of Griffith (Christensen 2007), but it suggested to the
author that perhaps the criterion could be developed
into something akin to the Hoek-Brown criterion.

323
ESTIMATING ROCK MASS PROPERTIES

Re-writing the Christensen criterion (Equation D3) in


terms of the octahedral stresses, which are:

or

The following equation is obtained.

EQUATION D5

Replacing the two terms in parentheses in Equation


D5 with parameters say, m1 and s1:

EQUATION D6

Where

and

324
14 | References

At the ductile limit, |σt| = σc, then m1 =0 and s1 =


2/9. At the brittle end of the spectrum and with |σt| →
0, then m1 → 2/3 and s1 → 0.
The similarity in form between Equation D6 and
the original Hoek-Brown criterion (Equation B2) can
clearly be seen.

EQUATION D7

So the constants b and c from Equation D4 can be


written in terms of m1 and s1 as follows.

In the same fashion as Hoek and Brown (1980) m1 and


s1 variables can be selected to best curve-fit test data.
The similarity between Equation D6 and the Hoek-
Brown criterion as extended to 3D by Pan and Hudson
(1988) (Equation D8) is also evident.

which can be re-arranged as

EQUATION D8

However, similar to the Pan and Hudson (1988)


equation as discussed in Zhang and Zhu (2007) and in
Appendix B, Equation D6 does not simplify to the 2D
Hoek-Brown criterion under triaxial stress state (σ2 = σ3).

Substituting σ2 = σ3 into Equation D6:

The term [3/2m1σ2σ1] on the right-hand side does


not appear in the Hoek-Brown criterion.

325

You might also like