Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

A.J.S. (Sam) Spearing, Liqiang Ma, Cong-An Ma - Mine Design, Planning and Sustainable Exploitation in The Digital Age-CRC Press (2023)

Download as pdf or txt
Download as pdf or txt
You are on page 1of 447

Mine Design, Planning and Sustainable

Exploitation in the Digital Age

Mine Design, Planning and Sustainable Exploitation in the Digital Age covers mine
planning, design and exploitation taking cognizance of new developments, especially
those associated with the Fourth Industrial Revolution and the positive influence that
it has, and will have, on the mining industry. It refers to latest best practices with
emphasis on the social license to operate and sustainable (green) mining.
The book covers surface and underground mining in some detail and addresses rel-
evant associated aspects such as risk management, green mining and the importance
of real community relations. It is organized as follows:

• Surface Mining
• Underground Soft Rock Mining
• Underground Hard Rock (Metal/Non-metal) Mining
• Green and Sustainable Mining

It has many relevant photos and figures that help the reader. It also includes appro-
priate support design and types commonly used in the various mining methods.
Mine Design, Planning and Sustainable Exploitation in the Digital Age is mainly
aimed at mining, geological engineering and other undergraduate and postgraduates
interested in the mining resources industry. It will also serve as a useful reference
book for practitioners in the mining industry who want an easy-to-use book.

A.J.S. (Sam) Spearing is a Professor in the School of Mines at the China University
of Mining and Technology (CUMT). Before that, he was the Director of the Western
Australian School of Mines in Kalgoorlie and is active in rock mechanics, mine design,
safety and support research. He has co-invented several support products commonly
used in tabular mining operations. He has been in academia for about 14 years and was
associated directly in the mining industry before that. He is the author of several books
and book chapters and has authored over 55 peer-reviewed papers.

Liqiang Ma is a Professor in the School of Mines at CUMT. He is the Deputy Dean of


the School of Mines and is active in water conservation mining, ground control, mine
design and rock mechanics research. He has been in academia for about 16 years. He is
the author of several books and book chapters and has authored over 100 peer-reviewed
papers. He spent a year at Penn State University as a visiting scholar from 2015 to
2016. His main research work is associated with mining methods (especially water con-
servation mining and ground control, including ground-breaking research in rockburst
precursors. He was selected as “Outstanding Talents of the New Century” by the
Chinese Ministry of Education

Cong-An Ma is an Associate Professor in the School of Mines at CUMT. Before that,


he was a Senior Engineer at the ChangSha Institute of Mining Research. He is active in
hard rock mining methods, mine rehabilitation and rock pass control theory research
for the sublevel caving mining method. He has been in academia for about 12 years
and was associated with the copper mining industry before that. He is the author of
two books and book chapters and has authored about 30 peer-reviewed papers. He
spent a year at the University of California Riverside as a visiting scholar from 2014
to 2015.
Mine Design, Planning and Sustainable
Exploitation in the Digital Age

A.J.S. (Sam) Spearing, Liqiang Ma, and


Cong-An Ma
School of Mines, China University of Mining
& Technology, Xuzhou, Jiangsu Province,
People’s Republic of China
Cover image: Photo courtesy of Dyno Nobel
First published 2023
by CRC Press/Balkema
Schipholweg 107C, 2316 XC Leiden, The Netherlands
e-mail: enquiries@taylorandfrancis.com
www.routledge.com – www.taylorandfrancis.com
CRC Press/Balkema is an imprint of the Taylor & Francis Group, an informa business
© 2023 A.J.S. (Sam) Spearing, Liqiang Ma, and Cong-An Ma
The right of A.J.S. (Sam) Spearing, Liqiang Ma, and Cong-An Ma to be identified as
authors of this work has been asserted in accordance with sections 77 and 78 of the
Copyright, Designs and Patents Act 1988.
All rights reserved. No part of this book may be reprinted or reproduced or utilised
in any form or by any electronic, mechanical, or other means, now known or hereafter
invented, including photocopying and recording, or in any information storage or
retrieval system, without permission in writing from the publishers.
Although all care is taken to ensure integrity and the quality of this publication and
the information herein, no responsibility is assumed by the publishers nor the author
for any damage to the property or persons as a result of operation or use of this
publication and/or the information contained herein.
British Library Cataloguing-in-Publication Data
A catalogue record for this book is available from the British Library
Library of Congress Cataloging-in-Publication Data
Names: A.J.S. (Sam) Spearing, author. | Liqiang Ma. author. | Cong-An Ma, author.
Title: Mine design, planning and sustainable exploitation in the digital age /
A. J. S. (Sam) Spearing, Liqiang Ma, and Cong-An Ma.
Description: Boca Raton: CRC Press, 2023. | Includes bibliographical references and
index.
Identifiers: LCCN 2022008826 (print) | LCCN 2022008827 (ebook)
Subjects: LCSH: Mining engineering—Technological innovations. | Sustainable engineering.
Classification: LCC TN145 .M25 2023 (print) | LCC TN145 (ebook) | DDC 622—dc23/
eng/20220627
LC record available at https://lccn.loc.gov/2022008826
LC ebook record available at https://lccn.loc.gov/2022008827
ISBN: 978-1-032-02873-6 (hbk)
ISBN: 978-1-032-02894-1 (pbk)
ISBN: 978-1-003-18568-0 (ebk)

DOI: 10.1201/9781003185680

Typeset in Times New Roman


by KnowledgeWorks Global Ltd.
Contents

1 Surface mining 1
1.1 Introduction and relevant terminology 1
1.1.1 Introduction 1
1.1.2 Relevant mining terminology 4
1.2 Strip mining and quarrying 13
1.2.1 Typical preparation and layout 13
1.2.1.1 Area mining 13
1.2.1.2 Quarrying 15
1.2.1.3 Contour mining 20
1.2.1.4 Mountaintop mining 21
1.2.2 Planning and operation 23
1.2.2.1 Drilling and blasting 24
1.2.3 Reclamation and rehabilitation 27
1.3 Openpit mining 28
1.3.1 Typical preparation and layout 28
1.3.1.1 Overall wall slope angle design 32
1.3.1.2 Bench height and angle design 32
1.3.1.3 Roadway design 34
1.3.2 Planning and operation 35
1.3.2.1 Openpit planning 36
1.3.2.2 Blasting 37
1.3.2.3 Dust 37
1.3.2.4 Stability monitoring 38
1.3.3 Rehabilitation/reclamation 38
1.4 Geotechnical design 39
1.4.1 Typical preparation and layout 39
1.4.1.1 Exploration 41
1.4.1.2 The final pit design 41
1.4.2 Geotechnical design processes 41
1.4.3 Geotechnical models 46
1.4.3.1 Experiential design 46
1.4.3.2 Empirical design 46
vi Contents

1.4.3.3 Mathematical design 46


1.4.3.4 Computer design 51
1.4.4 Geotechnical bench and slope design analysis
techniques 52
1.4.5 Slope monitoring 53
References 54

2 Underground soft rock mining 56


2.1 Introduction to soft rock mining 56
2.1.1 Coal formation and geology 56
2.1.2 Coal classification 58
2.1.2.1 Peat 59
2.1.2.2 Lignite 60
2.1.2.3 Sub-bituminous and bituminous coal 60
2.1.2.4 Anthracite 61
2.1.3 Coal exploration 63
2.1.4 Typical soft rock mining 64
2.2 Basic coal (soft rock) mine design and planning 66
2.2.1 Coal mine planning and design principles 66
2.2.2 Coal mine rockmass classification 67
2.2.2.1 Data collection for calculation of CMRR 67
2.2.2.2 The CMRR calculation process 68
2.2.2.3 Parameters involved in the calculation of
CMRR 68
2.2.2.4 Calculation of CMRR 71
2.2.3 Coal mine planning 74
2.2.3.1 The geological and resource model 75
2.2.3.2 The planning (engineering) model 75
2.2.3.3 The scheduling model 77
2.2.4 Mine planning challenges 79
2.3 Coal mine access and development 80
2.3.1 Introduction to coal mine access 80
2.3.2 Ramp (decline) access 80
2.3.3 Vertical shaft access 81
2.3.3.1 Shafts developed through overburden soil 82
2.3.3.2 Shafts in rock 84
2.3.4 Shallow shaft protection pillar design 89
2.3.5 Main entry development 89
2.3.6 Main access protection 91
2.4 Room and pillar mine design 91
2.4.1 Pillar design considerations 91
2.4.1.1 Main uses for pillars 91
2.4.1.2 Basic pillar design principles 92
2.4.1.3 Pillar load 95
Contents vii

2.4.2 Room and pillar panel design 96


2.4.2.1 Tributary Area Theory 96
2.4.2.2 Pillar strength 97
2.4.2.3 Pillar factors of safety 98
2.4.3 Barrier pillar design 100
2.5 Longwall mining 101
2.5.1 Longwall panel design 101
2.5.2 Stresses around a longwall panel 104
2.5.3 Longwall equipment 104
2.5.3.1 The Armoured Face Conveyor 106
2.5.3.2 The longwall plough (plow) and shearer 107
2.5.3.3 The longwall shields 110
2.5.4 Longwall entry (gate) design 113
2.5.4.1 Chain pillar design 113
2.5.5 The selection of design of longwall face supports 117
2.5.5.1 Determination of size of the support 117
2.5.5.2 Determination of roof loading on the support 118
2.5.5.3 Floor pressure under the base plate
of a longwall shield 118
2.5.6 Longwall subsidence 120
2.5.6.1 General overview of surface subsidence
associated with longwall coal mining 120
2.5.6.2 Estimating longwall surface subsidence 121
2.5.6.3 The surface subsidence zone 122
2.5.6.4 Some relevant definitions 122
2.5.6.5 Sub-critical panel 129
2.5.6.6 Critical panel 129
2.5.6.7 Super-critical panel 130
2.5.6.8 Sub-surface displacements and strains 130
2.6 Specialized mining methods 130
2.6.1 Coal mining with backfill 130
2.6.1.1 Backfilling a room and pillar coal mine 131
2.6.1.2 Backfilling a longwall coal mine 134
2.6.2 Top coaling 139
2.6.3 Ultra-high seams with single cut longwalling 141
2.6.4 Multi-cut coal mining 142
2.6.4.1 Descending slices 142
2.6.4.2 Ascending slices 143
2.7 Support design and application 144
2.7.1 Coal entry and room support design 144
2.7.1.1 Support challenges and requirements 144
2.7.1.2 Support design methodologies 146
2.7.2 Intersection support 155
2.7.3 Types of coal entry support 156
viii Contents

2.7.4 Longwall entry (gate) support 157


2.7.5 Longwall shield support design 159
2.7.6 Coal bump precautions 160
2.8 Coal mine hazards, safety and risk 161
2.8.1 Hazard identification and risk management 161
2.8.1.1 Hazard identification 161
2.8.1.2 Risk management in coal mines 162
2.8.2 Coal mine safety training 164
2.8.3 Coal mine hazard and safety monitoring 164
2.9 Coal mine rehabilitation and reclamation 165
2.9.1 Mine access closure 166
2.9.2 Coal mine waste rehabilitation 167
2.9.2.1 Avoiding coal mine waste 168
2.9.2.2 Closure and remediation of coal
mine waste 169
2.10 Green mining developments 172
2.10.1 Water conservation and subsidence control 172
2.10.1.1 The effects of subsidence and
the causal factors 173
2.10.1.2 Subsidence prediction 174
2.10.1.3 Coal mining subsidence 175
2.10.1.4 Potential hydrologic impacts of
underground mining 175
2.10.1.5 Impacts of mining-induced subsidence
on surface 179
2.10.2 Repurposing coal and thermal waste 179
2.10.3 Coal rockburst precursors, mitigation and control 180
2.10.3.1 Factors that can cause coal bursts 180
2.10.3.2 Coal burst precursors 181
2.10.3.3 Coal burst mitigation and control 181
References 181

3 Underground hard rock (metal/non-metal) mining 187


3.1 Introduction to hard rock mining 187
3.1.1 Typical hard rock geology 187
3.1.2 Hard rock orebody formation and classification 187
3.1.2.1 Igneous intrusions 188
3.1.2.2 Sedimentary concentration by weathering
and/or precipitation 188
3.1.2.3 Sea water extraction 188
3.1.2.4 Placer deposits (beaches and streams) 188
3.1.2.5 Pegmatite deposits (igneous deposits caused
by slow crystallization at high temperatures
and pressures) 188
Contents ix

3.1.2.6 Evaporite deposits 189


3.1.2.7 Chemically deposited sedimentary deposits 189
3.1.2.8 Other sedimentary deposits 189
3.1.2.9 Volcanic deposits 189
3.1.2.10 Metamorphic deposits 190
3.1.3 Hard rock exploration 190
3.1.3.1 Introduction to exploration 190
3.1.3.2 The Stages of Exploration 191
3.1.3.3 More details on exploration methods 193
3.1.3.4 Diamond (core) drilling 194
3.1.3.5 Rock sampling 196
3.1.4 Hard rock exploitation 196
3.2 Basic hard rock mine design and planning 197
3.2.1 Hard rock mine design principles 197
3.2.1.1 Rock Quality Designation (RQD) 199
3.2.1.2 Q System 200
3.2.1.3 Rock Mass Rating (RMR) 206
3.2.1.4 Mine Rock Mass Rating (MRMR) 209
3.2.2 The production cycle 213
3.2.3 Hard rock mine planning 214
3.2.4 Mine planning challenges and uncertainties 217
3.3 Mine access and development 218
3.3.1 Introduction to shaft and decline access 218
3.3.2 Shafts, raises, adits and ore (rock) passes 220
3.3.2.1 Shafts 220
3.3.2.2 Raises 222
3.3.2.3 Adits 225
3.3.3 Main access development 226
3.3.4 Main access protection 229
3.3.4.1 Shaft protection in a massive orebody 230
3.3.4.2 Shaft protection in a tabular orebody 230
3.4 Mine method selection 230
3.4.1 Mine design considerations and methodology 230
3.4.2 Rockmass classification 233
3.4.3 Basic mining method options 233
3.5 Unsupported mining methods 233
3.5.1 Longwall hard rock tabular mining (based
largely on Spearing, 1995) 235
3.5.1.1 Introduction 235
3.5.1.2 A typical hard rock longwall mine 235
3.5.1.3 The main hazards associated with
hard rock longwall mining 237
3.5.2 Sub-level caving 242
3.5.3 Block caving (Cave Mining) 246
x Contents

3.6 Naturally supported mining methods 252


3.6.1 Hard rock room and pillar mining 253
3.6.2 Sub-level and open stoping (modified
from Hamrin, 1982 & Queens University, 2011b) 256
3.7 Artificially supported mining methods 262
3.7.1 Drift and fill mining 262
3.7.2 Cut and fill mining 263
3.7.2.1 Overhand stoping 265
3.7.2.2 Up-hole stoping 267
3.7.2.3 In-front stoping 267
3.7.2.4 Backfill 267
3.7.2.5 Men and materials access way 269
3.7.3 Undercut and fill mining 271
3.7.3.1 The selection between cut and fill
and undercut and fill 271
3.7.4 Shrinkage mining 274
3.7.5 Vertical Crater Retreat (VCR) 278
3.8 Drilling and blasting 281
3.8.1 Introduction to blasting (www.science.jrank.org/
pages/2634/Explosives-History.html) 281
3.8.1.1 The history of explosives 281
3.8.1.2 Definition and components of an explosive 283
3.8.1.3 How explosives work and rock fragmentation 284
3.8.2 Optimizing blasting 286
3.8.2.1 Blasting in coal mines (and other fiery mines) 287
3.8.2.2 Blasting principles 287
3.8.2.3 Explosive Properties 288
3.8.3 Drilling and blasting equipment 292
3.8.3.1 Important requirements for rock drilling 292
3.8.3.2 Types of rock drills commonly used 294
3.8.4 Explosives and accessories 296
3.8.4.1 Main underground explosive types 296
3.8.4.2 Initiators and accessories 298
3.8.5 Underground blasting 298
3.8.5.1 Controlled blasting techniques 300
3.8.5.2 Tunnel excavation by blasting 301
3.8.5.3 The effect of high stresses on blast fracturing 302
3.8.6 Blast monitoring 304
3.9 Support design and application 305
3.9.1 Hardrock support design 305
3.9.1.1 The need for support and it’s functions 305
3.9.1.2 Support design overview 308
3.9.1.3 Design approaches 311
3.9.1.4 A successful support design 313
Contents xi

3.9.2 Rockbolts and cable bolts 313


3.9.2.1 Mechanical expansion shell 313
3.9.2.2 Deformed bar/rebar/ripple bar 314
3.9.2.3 Friction bolts (split sets) 315
3.9.2.4 Swellex bolts (Atlas Copco) 316
3.9.2.5 Resin-grouted bolts 316
3.9.2.6 Cable bolts 319
3.9.3 Specialized support 320
3.9.3.1 Yielding rockbolts 320
3.9.3.2 Self-drilling anchors 322
3.9.3.3 Spiling bolts 323
3.9.4 Shotcrete 323
3.9.4.1 Shotcrete history 323
3.9.4.2 The dry shotcrete machine and process 325
3.9.4.3 The wet shotcrete process and pump 327
3.9.4.4 Selecting the optimum shotcrete process
for the specific application 329
3.9.4.5 Admixture use 329
3.9.4.6 Shotcrete nozzles 330
3.9.4.7 Shotcrete production (in batch plants) 331
3.9.4.8 Shotcrete transportation from plant
to application site 331
3.9.4.9 Shotcrete costs 332
3.9.5 Backfill 334
3.9.5.1 Introduction to backfilling 334
3.9.5.2 Mining methods using backfill 335
3.9.5.3 Backfill transportation and design 337
3.9.5.4 Backfill criteria 339
3.9.5.5 Backfill placement 339
References 339

4 Green and Sustainable Mining 343


4.1 Green mining 343
4.1.1 The continued need for mining 343
4.1.2 United Nations Sustainable Development Goals 343
4.1.3 The need for green mining 344
4.1.4 The concepts of green mining 347
4.1.5 The Fourth Industrial Revolution and its
impact on mining 350
4.2 The circular mining economy 351
4.2.1 The current mining industry and the need for change 351
4.2.2 The need for a circular mining economy 352
4.2.3 Achieving a circular economy 355
4.2.4 The challenges to creating a circular mining economy 356
xii Contents

4.3 Sustainable development of minerals 357


4.3.1 Concepts and key factors 357
4.3.2 Environmental impact assessments 358
4.3.2.1 Introduction 358
4.3.2.2 The impact of mining 359
4.3.3 Mining sustainability and community relations 360
4.3.3.1 Mining sustainability 360
4.3.3.2 Community relations 361
4.3.4 Social license to operate 362
4.4 Mine waste disposal and repurposing 364
4.4.1 Waste dam design and disposal 364
4.4.1.1 Introduction 364
4.4.1.2 Engineering considerations for
tailings dam design 366
4.4.1.3 Conventional dam construction methods 367
4.4.2 Waste dam operation and monitoring 370
4.4.2.1 Dam operation 370
4.4.2.2 Main causes for tailings dam failures 372
4.4.2.3 Tailings dam monitoring 372
4.4.3 Tailings dam safety, design and operation
in the future 373
4.4.3.1 Tailings major hazards 373
4.4.3.2 Tailings dam future operations 374
4.4.4 Waste dam rehabilitation 376
4.4.5 Repurposing options and utilization 376
4.4.6 Challenges to repurpose tailings cost effectively 377
4.5 Urban mining 377
4.5.1 Introduction to urban mining 377
4.5.2 Waste suitability for urban mining 378
4.5.2.1 Plastics, paper and glass 378
4.5.2.2 Building and construction debris 382
4.5.2.3 Electronic waste 382
4.5.2.4 Battery waste 383
4.5.3 Challenges to growing urban mining 384
4.6 Mining risk and risk management 385
4.6.1 Introduction to risk 385
4.6.2 Mining hazards 386
4.6.2.1 Surface mining hazards 386
4.6.2.2 Underground mining hazards 387
4.6.2.3 Tolerable risk 387
4.6.2.4 Overcoming the fear of change 388
4.6.2.5 Identifying risks 388
4.6.3 Risk management 388
4.6.3.1 Managing risk 390
Contents xiii

4.6.3.2 Risky behaviour 390


4.6.3.3 Managing risk 391
4.6.4 Future mining technologies 392
4.6.4.1 Introduction 392
4.6.4.2 A major mining company’s vision 395
4.6.4.3 Space mining 397
4.7 Mining valuation and financing 399
4.7.1 Mineral resources 399
4.7.2 Mineral reserves 400
4.7.2.1 Resource and reserve reporting 401
4.7.3 Mineral economics 402
4.7.3.1 Mineral property valuations 403
4.7.3.2 Cost approach and associated methodologies 405
4.7.3.3 Market approach and associated methodologies 407
4.7.3.4 Income approach and associated methodologies 412
4.7.4 The time value of money 413
4.7.5 Market risk (β) 415
4.7.6 Country or sovereign risk 417
4.7.7 Project risk 418
4.7.8 Valuing shares (scrip) 419
4.7.9 Mining financing options 419
4.7.10 Equity 422
4.7.11 Debt 423
References 425

Index 428
Taylor & Francis
Taylor & Francis Group
http://taylorandfrancis.com
Chapter 1

Surface mining

1.1 Introduction and relevant terminology

1.1.1 Introduction
The aim of selecting a specific mining method should be to maximize the safe and
cost-effective extraction of an orebody whilst minimizing the environmental impact
and effectively collaborating with the local community (social license to operate) so
they can benefit long term from the mining.
The selection of the mining method is the most important decision that needs to be
made and depends on many factors, especially the technical, economic and human ones.
There are generally more than one mining method option for a deposit, and the selection
must be carefully considered based mainly on safety, profitability and the environment.
The main advantages of surface mining over underground mining are as follows:

• Lower operating and capital costs


• Higher recovery
• Shorter development times before exploitation
• Few (if any) equipment size and capacity constraints
• Typically less labour required
• Typically safer operation
• Higher productivity
• No potential for dangerous gas build-up (if applicable)

The main disadvantages are as follows:

• Blast vibrations and noise


• Dust
• Larger land and environmental disruptions
• Can be affected by adverse climate conditions such as heavy rain, snow, wind or heat

Warner (1934) defined the factors to be considered in detail as follows, and these have
not really changed:

Technical factors

• Topography
• Transportation of ore, personnel and equipment

DOI: 10.1201/9781003185680-1
2 Surface mining

• Infrastructure location (shafts, workshop, plant, etc.)


• Ore storage
• Materials storage
• Waste disposal
• Water drainage
• General stratigraphy and hydrology
• Water source, quantity, quality and temperature
• Water table level
• Water-bearing strata
• Impervious strata
• Seasonal variations
• Presence of major structural features
• Presence of major surface water sources (e.g., rivers, lakes, wetlands, etc.)
• Orebody
• Horizontal area of commercial ore zone
• Width-to-length relationship
• Vertical extent
• Boundary regularity
• Dip and/or pitch
• Spatial relationship to other economic orebodies or mines
• Presence of economic off-shoots, etc.
• Presence of waste intrusions
• Presence of major geological features traversing the orebody
• Presence of abnormally wide or narrow orebody sections
• Ore
• Value per unit of the ore
• Ease of concentration
• Value distribution (uniform or selective mining and/or sorting required)
• Weathering (e.g., swelling and/or oxidation)
• Health hazard creation (e.g., gases, dust, etc.)
• Explosion or fire hazard creation (e.g., methane, hydrogen sulphide, etc.)
• Mineralogy and micro-structure (e.g., tendency for packing in stopes or rock
passes)
• Rock breaking
• Walls and capping
• Strength (support and subsidence)
• Reaction to stress
• Failure nature (coarse/slabs/blocks or fine, free running or sticky material)
• Regularity and dip
• The mining location
• Climate
• Temperature gradient and variation on surface
• Temperature gradient and variation underground
Surface mining 3

Economic factors

• Labour
• Availability
• Skills and training levels
• Whether organized
• Political stability
• Services
• Supply, reliability and cost or electricity
• Supply, reliability and cost of fuel, etc.
• Material/equipment availability
• Capital availability
• Scale of operation
• Time from start to production
• Local infrastructure
• Distance to towns and resources (schools, hospitals, etc.)
• Size of nearby towns
• Local quality of life
• Demand and price fluctuation of the commodity being mined
• The effect of the new mine on the global supply and demand of the material mined

Human factors
• Relationship of the orebody to surface and public/private land
The only new factors that have become important since Warner in 1934 are related to
human and environmental ones, including:

• Environmental protection and safe mine closure


• Working effectively with the local community from exploration to mine closure
• Mine sustainability
• Social license to operate
• Much more focus on mine safety
• More emphasis on diversity and gender equality in the workforce

Surface mines normally only extend to a depth of about 200m, below which it is usually
cheaper to extract the ore using an underground mine. There are, however, massive
orebodies that outcrop on or near surface where the economic depth for openpit mining
is much more.
According to Mining Technology (February 2020), the five deepest openpit mines
in the world are:

1 Bingham Canyon, Utah, US, operated by Rio Tinto, has a pit width of over
4,000m and a current depth of 1,200m.
2 Chuquicamata copper mine, operated by Codelco in Chile, has a pit length of
4,300m, a width of 3,000m and a current depth of 850m.
4 Surface mining

3 Escondida copper mine in the Atacama Desert of Chile, operated by BHP, has a
pit length of 3,900m, a width of 2,700m and a current depth of 645m.
4 Udachny diamond mine located in Siberian, Russia, has a current depth of 630m.
5 Muruntau gold mine in Uzbekistan has just reached a depth of 600m.

The cut-off depth depends on the economies of the two fundamental mining methods.
Surface costs are dominated by the ore-to-waste (stripping) ratio, which, in turn, will
depend on the orebody geometry, the amount of overburden to be removed, the safe
overall pitwall angle and the individual bench height relative to the bench wall steep-
ness. The latter item will mainly depend on the type of rock, the number of fractures,
the presence of water, the underground mining method selected, etc.

1.1.2 Relevant mining terminology


Terminology often varies from country to country, and the following are figures and
definitions of the most commonly accepted terminology used in English-speaking
countries. It is not exhaustive but includes many of those often used.
Figures 1.1–1.5 illustrate many of the commonly used mining terms and their
descriptions.
The following is an alphabetic list of commonly used mining terminology and their
descriptions. It should be noted that these can differ slightly in different English-
speaking countries. The terminology and descriptions are largely based on:

• ISRM Glossary of terms.


• Rockburst Terminology – South African Chamber of Mines Circular 28/82.
• RPM Global’s Glossary of Mining Terms.

Figure 1.1 Commonly used openpit layout and terminology (Somagani, 2016)
Surface mining 5

Figure 1.2 Strip mining layout and terminology (Westcott et al., 2009)

ABRASION: Abrasive wear is also known as cutting wear and involves the wearing par-
ticle moving across the surface in constant contact with the surface or impinging
the surface at angles less than 45°.
ABUTMENT: The areas of unmined rock at the edges of mining excavations that may
carry elevated loads resulting from redistributions of stress if un-failed, otherwise
the elevated stress may be moved further into the solid rock.
ADIT: A near horizontal excavation (tunnel) from the surface by which a mine is
entered. A blind horizontal opening into a mountain or from a highwall.
AGGREGATE: Crushed or naturally occurring rock/stone that is typically added to con-
crete, shotcrete or backfill.
ANGLE OF INTERNAL FRICTION: The angle (ϕ) between the axis of normal stress and the
tangent to the Mohr envelope at a point representing a given failure to stress con-
dition for a solid material.
ANGLE OF REPOSE: The maximum angle with respect to a horizontal plane that the sur-
face of a pile of a loose material will assume.
ANISOTROPY: Having different properties in different directions.
ANTICLINE: An upward fold or arch of ductile rock strata.
AQUIFER: A bed of rock strata that contains water. The presence of aquifers must be
taken into consideration when designing any mine or excavation as it could be a
major hazard.

Figure 1.3 Room and pillar layout and terminology


Source: Appendix F: Underground Coal Mining Methods-National Academies Press
6 Surface mining

Figure 1.4 Typical retreat longwall mining layout and terminology


Source: Appendix F: Underground Coal Mining Methods-National Academies Press

ASSAY: The determination of the quantity of a desired metal per unit mass of the mate-
rial (rock) containing it. The term assay is generally restricted to materials con-
taining precious metals.
ATTENUATION: Reduction in the amplitude of a wave as the distance from the source
of propagation increases.

Figure 1.5 Example of a metal/non-metal mine layout and terminology


Source: Anatomy of a mine-National University of Columbia
Surface mining 7

BACK: The roof or upper part in any underground mining excavation.


BACKBREAK: Rock broken beyond the planned excavation outline; it is past the limit of
the last row of blast holes.
BACKFILL: Waste material, usually metallurgical tailings or waste rock that is used to
fill underground voids formed by mining.
BANK CUBIC METRE: A measure of overburden removed in mining operations. Equivalent
to one cubic metre of material in its undisturbed condition.
BATTER SLOPE: The sections of rockmass between catch berms within pit walls – usu-
ally excavated to a specific inclination/angle from the horizontal.
BED: A stratum of coal or other sedimentary deposit between two bedding planes.
BEDROCK: The more or less continuous body of rock that underlies the overburden soils.
BENCH OR BANK HEIGHT: The vertical height of a bank measured between its highest
point (crest) and its toe at the digging level or bench.
BERM: A horizontal shelf or ledge built into an embankment or sloping wall of an open-
pit or quarry. Its function is to break the continuity of an otherwise long slope for
the purpose of strengthening and increasing the stability of the slope or to catch or
arrest slope slough material. A berm may also be used as a haulage road or serve as
a bench above which material is excavated from a bank or bench face.
BLASTIBILITY: The index value of the resistance of a rock formation to blasting.
BLEEDER (BLEEDER ENTRY): Special air courses developed and maintained as part of a coal
mine ventilation system. They are designed to continuously move air-methane mix-
tures emitted by the goaf or at the active face away from the active workings and
into mine-return air courses. They are also called exhaust ventilation laterals.
BRITTLE FRACTURE: The sudden failure associated with the complete loss of cohesion
across a plane.
BUCKLING: Instability of a column or plate under sufficiently high load due to sudden
deflection of the structure.
BULK (INCOMPRESSIBILITY) MODULUS: Ratio of the hydrostatic pressure to the volumetric
strain that it produces.
BUMP OR BURST: A violent dislocation of the mine workings that is attributed to severe
stresses in the rock surrounding the workings.
BURDEN: Distance between the explosive charge and the free surface in the direction
of throw.
CAGE: A conveyance in a mineshaft used for hoisting men, material and equipment.
CAVITATION: The formation and subsequent collapse of vapour bubbles that occurs when
a slurry pressure drops and then rises above the vapour pressure of water. Cavitation
can occur in pipes when the pressure is reduced and the velocity increases, or when
backfill is supplied intermittently. It can cause serious damage and erosion.
CEMENT: It is usually defined as a binder, glue or adhesive. In hardened concrete, it
forms the matrix that binds the aggregates together to form the strong and rigid
composite structure.
CLOSURE: The reduction in the dimensions of an opening.
COEFFICIENT OF FRICTION: A constant of proportionality () relating the normal stress and
the corresponding critical shear stress at which sliding starts between two surfaces.
COHESION: The shear resistance at zero normal stress (intrinsic shear strength in
rocks). The shear strength is imparted to the material through the bonding of
adjacent particles by physical or chemical means.
8 Surface mining

COMMINUTION: The breaking, crushing or grinding of coal, ore or rock.


COMPACTION: The reduction in specific volume of the placed backfill arising from
gravity or external stresses.
COMPRESSIVE STRESS: Normal stress tending to shorten the body in the direction in
which it acts.
CONSOLIDATION: The process by which water drains from saturated backfill by perco-
lation. This causes a reduction in the specific volume and an increase in the density.
CONTRACTION: The linear strain associated with a decrease in length of a body.
CONTROLLED BLASTING: All forms of blasting designed to preserve the integrity of the
remaining rocks (e.g., smooth blasting, pre-splitting and post-splitting).
CONVERGENCE: The shortening between two basically parallel surfaces (usually the
hangingwall/roof and the footwall/floor).
CREEP: Time-dependent deformation.
CROSSCUT: A tunnel driven between the entry and its parallel air course (return air-
way) for ventilation and access purposes. It is also a tunnel driven from one seam
to another through or across the intervening measures; sometimes called ‘crosscut
tunnel’ or ‘breakthrough’. In vein mining, an entry perpendicular to the vein.
Also known as cut through.
DAMPING: The reduction in the amplitude of vibration of a body or system due to dis-
sipation of energy internally or by radiation.
DECAY TIME: The interval of time required for a pulse to decay from its maximum value
to some specified fraction of that value.
DECOUPLING: The ratio of the radius of the blasthole to the radius of the charge. This
causes a reduction in the strain wave amplitude by increasing the space between
the charge and the blasthole wall.
DEFORMATION: A change in shape or size of a solid body.
DEPOSIT: A mineral deposit or ore deposit is used to designate a natural occurrence
of a useful mineral, or an ore, in sufficient extent and degree of concentration to
justify exploitation.
DETONATION: An extremely rapid and violent chemical reaction causing the produc-
tion of a relatively large volume of gas.
DILATANCY: The property of volume increase under loading.
DILATION (VOLUMETRIC STRAIN): The quotient of the change in volume and the original
volume of an element of the material under stress.
DILUTION: Waste material adjacent to the ore that is generally unintentionally mined
with the ore.
DIP: The angle at which a stratum or other planar feature is inclined from the horizontal.
DISCONTINUITY SURFACE: Any surface across which some property for a rockmass is
discontinuous (e.g., bedding planes, fractures, etc.).
DISPERSION: The phenomenon of varying speed of transmission waves depending on
the relevant frequency.
DISPLACEMENT: A change in position of a point in a material.
DISTORTION: The change in shape of a solid body.
DUCTILITY: The condition in which a material can sustain permanent deformation
without losing its ability to resist load.
EARTHQUAKE: A sudden and violent shaking of the ground, sometimes causing
great destruction, as a result of movements within the earth’s crust or volcanic
action.
Surface mining 9

ELASTICITY: The property of a material whereby it returns to its original form or con-
dition after an applied force is removed.
ELASTIC LIMIT: The point on a stress versus strain curve at which transition from elastic
to inelastic behaviour occurs.
FAILURE: Where the maximum strength of a material is exceeded by an applied load.
FATIGUE: Decrease in the strength of a material caused by repetitive loading.
FOLIATION: Alignment of minerals into parallel layers, which can form planes of weakness/
discontinuities in rocks.
FOOTWALL: The mass of rock beneath a discontinuity surface (in tabular mining, the
rock below the reef plane).
FORCE: An action that tries to move an object from a stationary position or to change
its rate of movement or its direction of movement.
FRACTURE: A mechanical discontinuity in the rock.
FRANGIBLE: A material is frangible (brittle) if through deformation, it tends to break into
fragments rather than deforming plastically and retaining its cohesion as a single object.
FRICTION: The force between two surfaces tending to resist relative motion. It is caused
by the roughness of the surfaces of the bodies and acts tangentially between the
surfaces.
GAP GRADED: Backfill with a solids particle size distribution that has a deficiency of
particles in a specific size range.
GEOTECHNICAL DOMAIN: It is a volume of rock with generally similar geotechnical
rockmass properties.
GOAF (GOB): The part of a mine from which the mineral/ore has been partially or
wholly removed; the waste left in old workings. In a longwall coal mine, it refers
to the collapse of the roof behind the longwall shields.
GRAVITY: The force that tends to accelerate any object towards the centre of the earth
at a constant rate (9.81m/s2).
HANGINGWALL: The mass of rock above a discontinuity surface (in tabular mining, the
rock above the reef plane).
HAULAGE: A tunnel that is near horizontal used to transport men, material, equipment
and ore. It also is used to transport fresh air (ventilation).
HETEROGENEITY: Having different properties at different points within a material.
HIGHWALL: The unexcavated face of exposed overburden in a surface mine or in a face
(bank) on the uphill side of a contour mine excavation.
HOMOGENEITY: Having the same properties at all points within a material.
HYSTERESIS: The incomplete recovery of strain during unloading due to energy consumption.
INCLINE: Any entry to a mine that is not vertical (shaft) or horizontal (adit). Often incline
is reserved for entries that are too steep for a belt conveyor (+17° through to −18°),
in which case a hoist and guide rails are employed. A belt conveyor incline is
termed a slope. It is also a secondary inclined opening, driven upward to connect
levels, sometimes on the dip of a deposit; also called ‘inclined shaft’.
INDUCED STRESS: The stress that is due to the presence of an excavation. The magnitude
of the induced stress depends mainly on the level of the in-situ stress and the shape
and size of the excavation.
INELASTIC DEFORMATION: The portion of deformation under stress that is not annulled
by the removal of the stress.
IN SITU: It means in the natural or original position. Applied to rock or soils when
occurring in the location in which they were originally formed or deposited.
10 Surface mining

INTERBURDEN: The strata between two working (mineable) seams or deposits.


ISOTROPY: Having the same properties in all directions.
LIQUEFACTION: The phenomenon whereby the shear resistance of a placed backfill
mass decreases when subjected to loading and as a result the mass begins to mobi-
lize and flow.
LITHOLOGY: The character of a rock described in terms of its structure, colour, mineral
composition, grain size and arrangement of its component parts; all those visible
features that in the aggregate impart individuality of the rock.
MINERAL: A compound occurring naturally in the earth’s crust, with a distinctive set of
physical properties and a definite chemical composition.
MUCKING: The process of removing broken rock (usually after blasting).
MUCKPILE: The broken material at a developing face after being blasting.
NORMAL FORCE: A force directed normal (perpendicular) to the surface element across
which it acts.
NORMAL STRESS: The component of stress normal to the plane on which it acts.
ORE: Rock containing minerals that have economic value.
OVERBREAK: The quantity of rock that is removed beyond the planned perimeter of the
final excavation.
OVERBURDEN: Layers of soil and rock above a mineral deposit.
PERCOLATION: The movement of water, under hydrostatic pressure through the small
interstices of rock or soil.
PERMANENT (RESIDUAL) STRAIN: The strain remaining in a solid with respect to its ini-
tial condition after the application and removal of a stress greater than the yield
stress.
PERMEABILITY: The capacity of a material to allow the passage of liquid or gas.
PILLAR: An area of ground (usually ore) left within an underground mine to support
the overlying rockmass or hangingwall (roof).
PLANE STRESS (OR STRAIN): This is a state of stress (or strain) within a body in which all
the stress (or strain) components normal to a certain plane are zero.
PLASTICITY: Property of a material to continue to deform indefinitely whilst sustaining
a constant stress.
POISSON’S RATIO: The ratio of the shortening in the transverse direction to the elonga-
tion in the direction of an applied force in a body under tension below the pro-
portional limit.
POROSITY: The ratio of the aggregate volume of voids or interstices in a rock or soil to
its total volume.
PORTAL: The structure surrounding the immediate entrance to a mine; the mouth of
an adit or a tunnel.
PORTLAND CEMENT (PC or OPC): An inter-ground or separately ground mixture of
clinker and gypsum, with generally not more than 5% (by mass) of an inorganic
mineral addition. The particle relative density is about 3.14 and the loose bulk rel-
ative density is about 1.4. The typical specific surface (an indication of the fineness
of grind of the product) is around 300 m2/kg. As with all cement types, the exact
specification varies from country to country.
POTENTIAL ENERGY: The work a body is capable of doing when moving from one posi-
tion to another relative to the gravitational effect of the earth.
POZZOLAN: A siliceous or siliceous and aluminous material which in itself possesses
little or no cementitious value, but will, in finely divided form, and in the presence
Surface mining 11

of moisture, react with calcium hydroxide to form compounds possessing cemen-


titious properties.
PRIMITIVE (VIRGIN) STRESS: The state of stress in a geological formation before it is dis-
turbed by manmade operations.
PRINCIPAL STRESS (OR STRAIN): The stress (or strain) normal to one of three mutually
perpendicular planes on which the shear stress (or strains) at the point in the body
are zero.
RAISE: An inclined opening, vertical or near-vertical opening driven upward from a
level to connect with the level above, or sometimes even the surface.
RAPID-HARDENING PORTLAND CEMENT (RHPC): This has the same typical chemical com-
position of Portland cement but is ground finer, having a specific surface (area) of
typically around 400–500 m2/kg.
RESERVE: Probable reserve is established after further drilling and investigation of the
orebody (after a measured resource was established) which has raised the level
of confidence such that seeking initial project funding can usually commence.
Proven reserves are when the orebody is well understood, and the tonnages and
grades established beyond reasonable doubt.
RESOURCE: Indicated Resource is after initial drilling that has identified that there is
mineralization, but the configuration (orebody) is uncertain. A measured resource
is when the preliminary tonnage has been calculated but drilling is not sufficient
to be certain of the orebody’s continuity.
RIB: The side (walls) of a pillar or an excavation.
ROCKMASS: The sum total of the rock as it exists in place, taking into account the intact
rock material, groundwater, as well as joints, faults and other natural planes of
weakness that can divide the rock into interlocking blocks of varying sizes and
shapes.
ROCK MECHANICS: The theoretical and applied science of the mechanical behaviour of
rock.
ROCK: Any natural formed aggregate of mineral matter occurring in large masses or
fragments.
ROCKBURST: A seismic event that causes damage to underground workings.
ROCKFALL (FALL OF GROUND): A fall of a rock fragment or a portion of the fractured
rockmass without the simultaneous occurrence of a seismic event.
RUN OF MINE: Raw mined ore or material as it is transported to the plant (before
processing).
SCALING: The removal of loose rock from the roof or walls (ribs). This work is danger-
ous if done manually and a long bar (called a scaling bar) is often used.
SECANT (YOUNG’S) MODULUS: It is the slope of a line drawn from the origin of the stress-
strain diagram and intersecting the curve at the point of interest. Therefore, the
secant modulus can take different values depending on the location of intersect.
It is generally taken from the origin (start) to the end of the elastic limit in the
stress:strain test.
SEISMIC EVENT: A transient earth motion caused by a sudden release of the strain energy
stored in the rock.
SHAFT: A primary vertical or non-vertical opening through mine strata used for venti-
lation or drainage and/or for hoisting of personnel or materials; connects the sur-
face with underground workings. Shafts can also connect sub-surface workings
from one level to another.
12 Surface mining

SHEAR: A mode of failure where two pieces of rock tend to slide past each other. The
interface of the two surfaces of failed rock may represent a plane of weakness, or
a line of fracture through intact rock.
SLOPE: Any continuous face of rockmass within the overall pit wall (without stepping/
berms).
SPALLING: The longitudinal splitting in uniaxial compression or the breaking-off of
plate-like pieces from a free rock surface.
SPITTING: The violent ejection of splinters of rock from the surface of an excavation.
SPRINGLINE: The meeting of the roof arch and the sides of a tunnel. In a circular tunnel,
the springlines are at the opposite ends of the horizontal centreline.
STIFFNESS: The force versus the displacement ratio.
STRIKE: The bearing of a horizontal line in a plane or a joint.
STOPE: An extraction area that mines the mineral/ore, generally supported by sur-
rounding pillars of standing rock or waste walls in hard rock mining.
STRAIN BURST: Rockbursts at the lower end of the spectrum of violent events occurring
essentially at the surface of an excavation.
STRESS: The force acting across a surface element divided by the area of the element.
STRESS (OR STRAIN) FIELD: The ensemble of stress (or strain) states defined at all points
of an elastic solid.
STRESS (OR STRAIN) RATE: The rate of change of stress (or strain) with time.
STRENGTH: The maximum stress that a material can resist without failing for any given
loading regime.
STRESS RELAXATION: Stress release due to creep.
STRIKE: The direction of azimuth of a horizontal line in the plane of an inclined stra-
tum (or other planar feature) within a rockmass.
SUBSIDENCE: The downward movement of the overburden (soil and/or rock) lying
above an underground excavation or adjoining a surface excavation.
SUPPORT: A structure or structural feature built into or around an underground exca-
vation for maintaining its stability.
SWELLING: The constitutive mineralogy of the rock whereby water is absorbed caus-
ing a measurable volume increase. Swelling can exert very large time-dependent
forces on rock or support systems and reduce the size of excavations.
SYNCLINE: A fold in ductile rock in which the strata dips inward from both sides
towards the axis. The opposite is an anticline.
TANGENT (YOUNG’S) MODULUS: Slope of the tangent to the stress versus strain curve at a
given stress value (generally a stress equal to half the compressive strength).
TECTONIC FORCES: Forces acting in the earth’s crust over very large areas to produce
high horizontal stresses which cause can earthquakes. Tectonic forces are associ-
ated with the rock deforming processes in the earth’s crust.
TENSILE STRESS: The normal stress tending to lengthen the body along the direction in
which it acts.
THICKNESS: The perpendicular distance between bounding surfaces (e.g., bedding planes).
TRANSVERSE (SHEAR) WAVE: A wave in which the displacement at each point of the
medium is parallel to the wave front.
TRIAXIAL COMPRESSION: Compression caused by the application of normal stress in
three perpendicular (orthogonal) directions.
UNIAXIAL (UNCONFINED) COMPRESSION (UCS): Compression caused by the application of
normal stress in a single direction.
Surface mining 13

VECTOR: A quantity having magnitude and direction but no fixed position.


WAVEFRONT: A continuous surface over which the phase of a wave that progresses in
three dimensions is constant.
WEATHERING: The process of disintegration and decomposition as a consequence of
exposure to the atmosphere, to chemical action and to the action of frost, water
and heat.
WEIGHT: The force with which the earth attracts a body.
YIELD STRESS: The stress beyond which the induced deformation is not fully annulled
after complete distressing.
YOUNG’S MODULUS OR MODULUS OF ELASTICITY E: It is a tensile mechanical property that
measures the elastic stiffness of a solid. It gives the relations between the ten-
sile stress σ and axial strain ε. The relationship is related to Hook’s Law and
E=σ/ε. The magnitude of Young’s Modulus in rock is large and usually expressed
in gigapascals (GPa).

1.2 Strip mining and quarrying

1.2.1 Typical preparation and layout


There are two common strip mining methods that depend on the surface topography,
the orebody depth and geometry:

• Area mining: It is used for near surface relatively flat deposits (typically coal, tar
sands and phosphates). It operates typically by removing long parallel trenches or
strips that are progressively mined and rehabilitated after a strip is fully mined out.
• Contour mining: It follows the outcrop of the deposit typically in a hilly or moun-
tainous area.

The mining cycle for both methods is usually the same and consists of:

• Clearing of the vegetation


• Soil and sub-soil removal
• Drilling, blasting and removal of the rock overburden
• Drilling, blasting and removal of the orebody
• Reclamation (that consists of rock replacement, followed by sub-soil and finally
soil replacement and finally appropriate revegetation)

1.2.1.1 Area mining


In this method, the following are the steps:

• The first step is the removal of all the vegetation, which is normally done using
dozers or scrapers. Depending on the Environmental Impact Study, the vegeta-
tion may need to be replaced after the strip mining in which case a local plant
nursery may be established.
• The soil, and, if necessary, the sub-soil, is removed using scrapers and dozers. The
soil is dumped into a specific area so it can be replaced last during the rehabilita-
tion stage.
14 Surface mining

• If there is overburden rock above the deposit, this is drilled and blasted and
removed onto a spoil heap.
• The deposit is then drilled, blasted and removed.
• The rock is then moved back from the spoil pile, followed by the subsoil and the
soil. This is usually completed to the original topography before mining but often
lower as the deposit has been removed. The use of global positioned systems on
the earthmoving equipment and survey drones enables this process to be much
quicker and efficient than it used to be.
• Vegetation is replanted as per the mining regulations and hopefully with active
input from the local community.

The replaced topsoil supports the revegetation and holds the moisture needed for
growth. The removal of the vegetation and topsoil is the most critical operation as
after the strip mining, the topsoil is replaced last, and this is critical to the rehabilita-
tion especially the revegetation. Once removed, it is carefully stockpiled.
Often the loose materials are a combination of topsoil, subsoil, etc., referred to as
unclassified material and cannot be separated so it is removed and stockpiled “as is”.
The rock above the deposit is now exposed and then drilled, blasted and removed to
a rolling stockpile.
Figure 1.6 shows some of a typical strip mine stages.
Figure 1.7 shows a huge strip coal mine in Germany.
Depending on the scale of the strip mining, the clearing of the soil and overburden
can be undertaken using dozers and excavators or even massive draglines (if the over-
burden depth is very large).

Figure 1.6 The strip area mining steps


Source: Arch Coal Inc. (Dec 2008) Annual Report Pursuant to Section 13 or 15(d) of the SECURITIES
EXCHANGE ACT of 1934 Form 10K – United States Securities and Exchange Commission
Surface mining 15

Figure 1.7 The Jänschwalde huge strip mine


Reproduced with permission from Lusatian Energy Stock Company, Germany

1.2.1.2 Quarrying
This term is generally used for surface mining of dimension stone such as granite,
marble, slate and sandstone. These are produced as blocks of stone that are commonly
used as decorative facing on buildings, both internally and externally.
They are generally mined with vertical faces from a few metres to over 50-m high.
The mining is conducted typically using non-explosive methods (or low explosive
methods). The mining is therefore slow, highly selective and costly. The orientation of
the faces is dictated by local geology, taking account specifically of the direction of the
prominent minor faults and joints.
The production of the blocks involves freeing the block on all sides except one. The
block is then totally freed by cutting it loose from the rockmass typically using a saw
or impregnated wire. The blocks are then sold or resized into smaller blocks or slabs.
The mass of the blocks can be from 10t to 50t.
Examples of quarries are shown in Figure 1.8.
The size of the block that can be extracted is key to the economic viability of the
quarry. Understanding the joints and their orientation, condition, length and strength
is vital.
Production of the blocks is achieved by one or, more commonly, a combination of
the following:

• Wire cutting
• Sawing
• Drilling and wedging closely spaced holes
16 Surface mining

Figure 1.8 Examples of stone quarries. (a) Albion Mine and Quarry, Isle of Portland, Dorset.
(Reproduced with kind permission from Dr. David Giles, Card Geotechnics Ltd.) (b) One
of the largest marble quarries in the world. (Reproduced with kind permission from RK
Marble Mine, India.)

Figure 1.9 Diamond wire cutting dimension stone


Source: Almasi, S. & Khademian, A. (2015) Influence of cutting wire tension on travertine cutting rate.
24th International Mining Congress and Exhibition, Turkey. April
Surface mining 17

Wire cutting with diamond-impregnated wire is a popular way of preparing dimen-


sion stone blocks. The parameters that need to be considered are both controllable
and uncontrollable. The controllable parameters are associated with the equipment
type, power and wire selected and the uncontrollable ones are associated mainly with
the local geology and mineralogy of the deposit. To start a cut, the diamond wire
cables are threaded through drill holes.
Figure 1.10 shows a 10 mm × 5.5 mm electroplated diamond impregnated wire used
for cutting dimension stone.

Figure 1.10 An example of a diamond wire used for dimension stone cutting.
Reproduced with permission from Cuts Inc, Knoxville, TN, US
18 Surface mining

Figure 1.11 Sawing dimension stone


Reproduced with kind permission from Heldeberg Bluestone

More conventional saws can be used, but they are not as versatile as the diamond wire.
Minimizing production cost and wastage are the main factors determining the block extrac-
tion method. Figure 1.11 shows a saw being used to cut dimension stone and Figure 1.12
shows a block being cut with a saw at its base so the freed block can then be moved.
Where drilling and blasting are done, the holes are very closely spaced and a light
charge is used (pre-splitting) to limit the blast damage (as shown in Figure 1.13).

Figure 1.12 Forming a block by undercutting it with a saw


Reproduced with permission from T. Heldal, Norwegian Geological Survey (NGU) Feb 2020
Surface mining 19

Figure 1.13 A high productivity dimension stone drilling rig


Reproduced with permission from Sandvik
20 Surface mining

Figure 1.14 Breaking dimension stone using expansive grouts


Reproduced with kind permission from Dexpan, www.dexpan.com

Even micro-cracks caused by blasting can create additional waste product. Rather
than using explosives, expansive grouts can be used, which have obvious safety ben-
efits and creates less wastage but can be slower and less productive. This is shown
in Figure 1.14.
Generally, diamond wire is the most commonly used method.

1.2.1.3 Contour mining


In contour strip mining, strips are excavated around a hill or mountain generally fol-
lowing a contour where the seam/orebody is located (as shown in Figure 1.15). It allows
the partial removal of a seam (typically coal) in a hill or mountain where the complete
removal of the seam may not be economic or safe for example. Contour mining and
mountaintop removal are therefore related.
The method is generally as follows:

• Access up the hill or mountain is developed to the seam and around it where con-
tour mining is planned.
• Vegetation is removed from the site.
• A bench is excavated in the mountain at the coal elevation, to give safe and easy
access to the equipment required for mining.
• Drills are used to remove the overburden and expose the seam. The rock is loaded
with shovels or loaders and removed by trucks and frequently stored for later
backfilling to nearly recreate the previous contours.
• The ore (generally coal) is then mined in the same way using drilling and
blasting.
• Additional coal can also be extracted using some form of conventional highwall
mining using a continuous miner or auger drills.

The blasting and dust can be effectively controlled so as not to cause adverse local
problems, but the main issue is potential water pollution as waste and water from
Surface mining 21

Figure 1.15 Simplified sketch of contour mining


Source: Mining Practices & Impacts, D. Willard, April 2012 – Slideshare

the mining naturally ends up in the valleys and therefore into the streams due to
gravity.
Once mining is complete, the area is returned to a condition specified by the min-
ing permit and applicable mining regulations. The recovery of the vegetation can be
slow as it is more difficult to repair the steep contours associated with the hills and
mountains.
Historically, the waste rock has unfortunately sometimes been disposed of ran-
domly over the side of the hill and mountain and this causes serious environmental
damage.

1.2.1.4 Mountaintop mining


Mountaintop mining achieves basically complete extraction of the seam. It requires a
relatively small workforce, is cost effective but produces a large quantity of waste rock
(spoils) and the disposal of this can create significant environmental issues.
The overburden is most often disposed of in adjacent valleys as few other eco-
nomic alternatives are available. Spoil disposal areas (valley fills) are therefore created
downstream.
An example of mountaintop mining and the waste disposal method is shown in
Figure 1.16.
As noted in the US Environmental Protection Agency’s report titled: “The Effects
of Mountaintop Mines and Valley Fills on Aquatic Ecosystems of the Central
Appalachian Coalfields”, mountaintop mines and valley fills lead directly to five prin-
cipal alterations of stream ecosystems:

• Springs and ephemeral, intermittent and perennial streams are permanently lost
with the removal of the mountain and from burial under fill.
22 Surface mining

Figure 1.16 Contour mining (mountaintop removal)


Source: Environmental Health Perspectives, Vol 119, No 11
Surface mining 23

• Concentrations of major chemical ions are persistently elevated downstream,


degraded water quality reaches levels that are acutely lethal to organisms in
standard aquatic toxicity tests.
• Selenium concentrations are elevated, reaching concentrations that have caused
toxic effects in fish and birds, and macroinvertebrate and fish communities are
consistently degraded.
• The scientific literature suggests that the headwater stream resources lost under
valley fills may not be successfully reconstructed at the completion of reclamation
of surface/mountaintop mines.
• It can also lead to large-scale landscape changes: loss of large tracts of forested
areas, fragmentation of forests and conversion of habitats (i.e., forest to grass-
lands) resulting in displacement or loss of species.

1.2.2 Planning and operation


According to Schissler (2004):

The primary planning mechanism used in strip mining is the range diagram,
which is a cross-sectional plan of the shape of the pit in various stages of min-
ing. The range diagram allows the dragline or continuous excavator equipment
characteristics of dig depth, reach, and physical size to be placed on the geologic
dimensions of depth to seams (overburden), and depth between seams (interbur-
den). By comparing machinery specifications with dimensional characteristics of
the geology, the mine designer can plan the pit width and dig depth (Figure 1.17).
As the dragline or continuous excavator moves the overburden to the adjacent
empty pit where the coal has been removed, the rock swells in volume. Earth or

Figure 1.17 Example of a range diagram using a dragline.


Source: Carlson
24 Surface mining

rock increases in volume, called the swell factor, when the material is removed
from its in situ or in-ground state and placed into a pit or on the surface. The range
diagram allows the mine planner to identify the equipment dump height required
to keep the displaced overburden (spoil) from crowding the machinery and min-
ing operations. In certain cases of mining multiple coal seams from one pit, a coal
seam can provide the boundary between the pre-strip and strip elevations.

1.2.2.1 Drilling and blasting


The aim of drilling and blasting is to cost effectively break the rock (initially overbur-
den then the ore itself) to a size that can be easily and effectively loaded and trans-
ported using the equipment suite available.
The hole length drilled is influenced by:

• The overburden and orebody geometry or desired excavation profile


• Production needs or excavation targets
• Available drilling equipment
• Drilling accuracy

The hole diameter is influenced by the following factors:

• Desired production rate


• Bench height
• Desired fragmentation distribution
• Geological structure
• Rock hardness
• Explosive type
• Site sensitivity
• Excavation technique

The results from drilling and blasting must always be re-evaluated and continuously
improved and optimized as it has the most influence on production costs (and produc-
tivity) for any given orebody. The process is shown in Figure 1.18.

Figure 1.18 Optimum blast monitoring and performance


Source: Blasting Dynamics, Dyno Westfarmers Ltd.
Surface mining 25

Figure 1.19 Factors effecting explosive performance

Successful drilling and blasting depend on matching both with each other and con-
sidering mainly the rock properties and local geology. Figure 1.19 shows the three main
parameters that influence the performance of the actual explosive used in the blast hole.
There are basically four different blasting patterns commonly used on surface mines:

• Row by row
• Chevron
• Echelon
• Diamond

Row by row: This is used for bench blasting where a free face already exists. This fir-
ing pattern can be applied in a row-by-row initiation sequence with delay times only
between rows or in a pattern with short delay times between holes in a specific row and
long delay times between rows so there is no simultaneous blasting of the rows. This
design requires at least one free face. A typical layout is shown in Figure 1.20.
Modified row by row: This is the same basic method as row by row except that the
timing of the edge holes is delayed as shown below to try and better concentrate the
broken rock (muckpile) more centrally (as shown in Figure 1.21).

Figure 1.20 Row by row drill pattern


Source: Hagan & Mercer, 1983
26 Surface mining

Figure 1.21 Modified row by row and initiation sequence


Source: Dyno Nobel

Chevron: In this firing pattern, the delays between holes and rows are chosen so that
the firing sequence results in a V-formation. By using different delays, the angle of the
V-formation can be modified. The Chevron design also requires at least one free face
and produces a concentrated muckpile as shown in Figure 1.22.
Echelon: The Echelon firing pattern is simply one half of a Chevron pattern. The
echelon pattern however requires two free faces as shown in Figure 1.23.

Figure 1.22 Chevron blasting pattern and initiation sequence


Source: Hagan & Mercer, 1983

Figure 1.23 The echelon blasting pattern and initiation sequence


Source: Hagan & Mercer, 1983
Surface mining 27

Figure 1.24 The diamond drill pattern and initiation sequence


Source: Dyno Nobel

Diamond: Diamond firing pattern is used for box cuts, sump blasts and other
applications (such as overburden removal) where there are no free faces parallel
to the blast holes. The broken rock will be displaced upwards, therefore there
is a risk of f ly-rock. A typical blast hole and initiation sequence is shown in
Figure 1.24.
A bench blast sequence is shown in Figure 1.25.

1.2.3 Reclamation and rehabilitation


After mining finishes, the mined area must undergo effective rehabilitation. Waste
dumps are generally contoured to flatten them out, in order to stabilize them for the
long term. If the ore contains sulphides, it is usually covered with a layer of clay to
prevent access of rain and oxygen from the air, which can oxidize the sulphides to
produce sulphuric acid, a phenomenon known as acid mine drainage. This is then
generally covered with soil, and vegetation is planted to help consolidate the material.
Eventually this layer will erode, but it is generally hoped that the rate of leaching or
acid will be slowed by the cover such that the environment can handle the load of acid
and associated heavy metals. There are no long-term studies on the success of these
covers due to the relatively short time in which large-scale openpit mining has existed.
It may take hundreds to thousands of years for some waste dumps to become “acid
neutral” and stop leaching to the environment. The dumps are usually fenced off to
prevent livestock denuding them of vegetation, and keep the public away. The openpit
is then surrounded with a fence, to prevent access, and it generally eventually fills up
with ground water. In arid areas, it may not fill due to the deep groundwater levels and
high evaporation rates.
28 Surface mining

Figure 1.25 Bench blasting mining cycle

1.3 Openpit mining

1.3.1 Typical preparation and layout


Openpit mines are used when generally massive deposits of commercially useful min-
erals or rock are found near the surface, where the overburden (surface material cov-
ering the valuable deposit) is relatively thin or the material of interest is structurally
unsuitable for underground excavation (as would be the case for sand and gravel).
For minerals that occur deep below the surface, where the overburden is thick or the
mineral occurs as veins in hard rock, underground mining methods are more cost
effective to extract the desired material. Many massive deposits may start as openpit
mines but later proceed underground if the deposit continues deeper than can be eco-
nomically recovered from the surface. The underground operation should ideally be
started before the surface extraction ceases.
Openpits are used mainly for the following orebodies:

• Large massive orebodies


• Generally steeply dipping orebodies
• Orebodies relatively close to the surface or out-cropping on surface

Hartman and Mutmansky (2002) give openpit layout examples based on different ore-
bodies as shown in Figure 1.26.
Surface mining 29

Figure 1.26 Openpit layouts depending on the orebody

Surface mining is less costly, more productive and safer than underground mining
but it can cause potentially more adverse environmental impacts.
A typical large openpit mine is shown in Figure 1.27.
Calder et al. (1997) list the key drivers of an openpit operation as:

• Ore grade and tonnage


• Topography
• Physical size, shape and structure of the orebody
• Required capital expenditures
• Net present value (NPV) and costings
• Pit limits, cut-off grade and stripping ratio
• Mining equipment capacity and fleet
• Required (optimum) rate of production
• Access and general infrastructure
• Pit design (pit angle, bench height and angle, road grades, etc.)
• Geotechnical aspects including rockmass properties and local geology
• Hydrological conditions

Additional considerations include:

• The mining lease area


• Overall stripping ratio
30 Surface mining

Figure 1.27 The Superpit operated by Kalgoorlie Consolidated Gold Mines, Western Australia
Photo taken by A. Spearing from the public viewing platform

Typical equipment that is used includes:

• Hydraulic shovels
• Front end loaders
• Haul trucks
• Drill rigs
• Scrapers
• Dozers
• Excavators
• Explosive trucks
• Graders
• Water trucks

A typical open pit mine showing some important features and operations is shown on
Figure 1.28, and typical terminology is illustrated on Figure 1.29.
An openpit is mined to its ultimate pit limit using a series of “nesting” openpits that
start small and get progressively larger, and are not necessarily uniformly created. To
achieve this, the pit is mined in pushbacks (also called phases or cutbacks) that must
each be optimized so that the NPV of the orebody is maximized over the mining life.
During the mining life of the openpit mine, the ultimate pit limit can change often
based on changes to the value of the ore mined or changes in the rock properties.
Surface mining 31

Figure 1.28 An openpit mine showing important features and equipment


Reproduced with kind permission from Earth Science, Australia

Figure 1.29 Openpit terminology and the overall pit wall and bench angles (Arteaga et al., 2014)
32 Surface mining

Figure 1.30 An example of nesting openpits created by pushbacks (Songolo, 2010)

As an openpit starts, it is mined in a series of pushbacks that involves making the pit
wider and deeper, which involves the blasting of benches an example of which is shown
in Figure 1.30.
Stripping ratio (SR) is the tonnes of waste (often overburden) to the tonnes of ore
mined in any cut. There is generally an inverse relationship between the orebody grade
and the stripping ratio.

SR = Waste mined/Ore mined (both typically in tons)

As the openpit is developed, the stripping ratio changes depending mainly on the
extent and grade consistency of the orebody.
Maximum allowable stripping ratio is the stripping ratio at which the costs to
uncover and extract the ore equal the selling price of the ore produced. Uncovering
any deeper ore than this critical ratio would therefore result in a loss of money.
Maximum allowable stripping ratio (SRmax) is therefore defined as:
SR max = [(ore value) − (production cost)]/(stripping cost)]

1.3.1.1 Overall wall slope angle design


The overall wall slope angle is the angle measured from the bottom of the openpit
(at the bottom bench toe) to the top of the openpit (the top bench crest) as shown in
Figure 1.31 It is determined by:

• Local geology (especially the properties and orientation of faults, joints and beds)
• Hydrology
• Rockmass properties (including the weathering potential)
• Depth of the final pit

To reduce the amount of waste mined and hence increase the economic depth that can
be mined, the slope angle must be as steep as safely possible. It can also vary around
the pit depending on local rockmass conditions.

1.3.1.2 Bench height and angle design


Overall pit slopes are divided into smaller and steeper individual benches typically
between 5 and 20-m high (vertical height), which can be drilled and blasted more
Surface mining 33

Figure 1.31 Bench and pit stability (Somagani, 2016)

stably and reliably and any lose rock will be arrested on the flat berm at the bottom
of each bench and not fall further, causing more dangerous conditions towards the pit
bottom. Ideally all benches should be the same height to facilitate easier, faster and
more effective drilling.
The bench height depends on:

• Local rockmass properties, local geology and hydrology (of the ore and waste)
• Equipment availability
• Planned production rates
• Weather

Ideally the bench should be as steep and as high as safely possible to maximize
productivity.
The major hazards associated with benches are rockfalls. The consequences are
serious and can include: death, injury, equipment damage and loss and access loss.
Rockfalls are generally associated with geological features that may have been weaker
than anticipated or unknown during the design and planning process. Poor blasting
practice can also result in safety and stability issues. The important features and lay-
out of openpit benches and roadways is shown in Figure 1.31.
The stability hazards can be mitigated by considering:

• Local geology especially fault and major joint orientation


• Hydrology
• Rockmass properties
• The blast size
• Pre-splitting (smooth wall blasting)
• Scaling if accessible and if it can be safely undertaken
• Where possible, limiting access to benches
• Rock reinforcement as a last resort generally near major pit roadways if necessary
• Burden control especially of the toe
34 Surface mining

Figure 1.32 Common bench failure modes (Prajapati, 2016)

The typical failure modes of benches or the overall pit include:

• Toppling failure
• Wedge failure
• Circular failure (of soils)
• Plane failure (could be rotational failure)

These failure modes are shown in Figure 1.32

1.3.1.3 Roadway design


The optimum roadway design is critical to minimize production costs. The gradi-
ent, condition, width and camber are very important design parameters. It must be
remembered that typically fully loaded vehicles are travelling up-dip and generally
empty vehicles are travelling down-dip (back into the pit). The maximum road gra-
dient is generally 10°, with 6–8° being common. Steeper dips become uneconomical
mainly because the trucks are moving out of the pit fully loaded.
Most mines use local waste material for road construction as that is the most cost
effective and readily available.
Important roadway construction considerations include:

• Using strong and stable rock


• Having a good size distribution to maximize compaction
• Having the correct road gradient and camber to ensure good drainage

Roadway construction equipment requires the use of:

• Dozers
• Graders
• Compactors
Surface mining 35

Ongoing road maintenance is essential to try and minimize transport costs, as poor
conditions significantly reduce tire life, slow transport speeds, increase vehicle main-
tenance costs and downtime and also increase driver fatigue with possible associated
skeletal injuries. The most common maintenance issues are caused by rain, freeze-thaw,
spillage from vehicles and heavy traffic. Aside from regrading and repairing the road
surface, water trucks need to be regularly used to reduce dust and improve visibility.
The road width depends mainly on the size of the trucks and varies typically from
about 25–40m.
According to Tannant and Regensburg (2001), the following designs are common:

• Average ditch widths and depths were 3m and 1m respectively.


• The height of roadway safety berms where generally 50–75% of the largest vehicle
wheel diameter.

Further they list the roadway key planning and alignment issues as:

• Stopping distances – the road alignment must meet the length of the longest (and
loaded) vehicle’s stopping distance.
• Driver sight distances – these must be at least the length of the longest stopping
distance of the vehicles in the fleet and must consider road curves and turns.
• Road widths and cross slopes – the useful width of the road is usually three to four
times the width of the largest vehicle with curves and turns usually even wider.
Curves are usually cambered at 1:25 for drainage and vehicle stability considerations.
• Curves and super-elevation – Horizontal curves should be designed to ensure that
all the vehicles can safely negotiate the curve at a given maximum speed, considering
sight distances and safe minimum turning radius. The super-elevation of the curve
is required to reduce the centrifugal forces on the truck when it negotiates the curve.
• Super-elevation run-out – when moving from a level road section to a super-
elevation curve, there must be a gradual and consistent increase so the vehicle can
safely navigate the curve.
• Maximum and sustained grade – typically these are kept at an average of 3.5–6°
(6–10%) but may be as much as 12° (20%) for a short length. A long constant grade
is more effective than a combination of shorter steeper and shallower gradients.
• Intersections – they should be made as flat as possible and not at the top of a ramp.

1.3.2 Planning and operation


Planning and operation of a mine depends on taking the geological data, modelling it,
determining target production levels, planning the mining then scheduling the opera-
tions. All this must also be financially justified.
Hartman (1992) details the following parameters for production scheduling of operations:

• Minimizing preproduction costs


• Assuring adequate working room
• Smoothing of the stripping ratios
• Timely exposure of ore grade material
• Reclamation accountability
• Maximizing production
36 Surface mining

The planning and modelling of various pushbacks allow the mine planner to opti-
mize the NPV of the orebody, which is critical.

1.3.2.1 Openpit planning


Mine planning must meet safety and environmental regulations and community
expectations, whilst maximizing the financial value of the asset (the orebody).
Camus (2002) has classified mine planning into two categories that are function
dependent and not time scale dependent:

• Strategic Planning is where the main decisions that govern orebody extrac-
tion are made, and the main goal is determining the overall financial returns
for the project. The decisions are related to the mining method selection,
infrastructure needs, mine planning, cut-off grade determination and cost
estimates.
• Tactical Planning is where the detail planning is undertaken. It includes the pro-
duction scheduling from the start-up-, short- and medium-term production plans,
budgeting, equipment and manpower schedule and needs.

The planning process starts by taking the geological data and model and transform-
ing it into a Block Model. Each Block has a certain volume and is assigned a certain
grade or value. The size of the Block is determined by considering mainly:

• The exploration borehole spacing


• The orebody geometry

The grade estimation for each Block is determined mathematically using various
methods such as Kriging and the economic value of each Block is determined, ignor-
ing mine access and development costs. The goal of openpit planning is to maximize
the NPV of the orebody, but this is easier “said than done”. Consider Whittle’s (1989)
quote:

The pit outline with the highest value cannot be determined until the block values
are known. The block values are not known until the mining sequence is deter-
mined; and the mining sequence cannot be determined unless the pit outline is
available.

The openpit planning process is therefore clearly complex and involves considering
many different iterative design comparisons using a calculated cut-off grade. This
involves developing a series of nested pits (pushbacks) from the initial one, which is
the most profitable as it has the least waste dilution to the final one (which is the pit
limit). The final pushback is generally the least profitable as it is the deepest one and
has the highest associated waste dilution (stripping ratio).
The long-term planning process, or life of mine planning, is a circular iterative pro-
cess as shown in Figure 1.33
The planning process is dynamic and must be continually reviewed and updated
based on new information, such as updated block values (from more drilling or sam-
pling for example) or change in the commodity price or working costs.
Surface mining 37

Figure 1.33 The long-term planning process (Dagdelen, 2000)

1.3.2.2 Blasting
Blasting is the key to the operation and hence must be done safely and effectively,
which requires ongoing monitoring to ensure optimal fragmentation and minimum
flyrock and dust. Flyrock can be caused by:

• Incorrect blast design


• Poor or no stemming and tamping
• Incorrect burden (hole spacing)
• Local geology
• Incorrect charging or timing

Safe blasting requires:

• An optimally designed blast for the local conditions and this can change as dic-
tated by changes in conditions
• Safe procedures to ensure that people and equipment are at a safe distance
• Security to keep unauthorized people off the site and around the blast site to
ensure no one is in danger
• Blasting signals that are well understood (to signal an imminent blast and later
an all-clear)

1.3.2.3 Dust
Dust is a potential health problem to the workforce and the local community and can
be caused by drilling and blasting, vehicle travel on unpaved roadways and from waste
dumps or dams.
The effects of dust can be mitigated by various strategies including:

• Optimum blasting
• Regular watering of unpaved roads or using a suitable polymer or chemical
treatment
• Providing vehicles with sealed and well-ventilated cabins with effective dust filters
38 Surface mining

1.3.2.4 Stability monitoring


Measuring the pit wall stability during mining is very important and has become a
very useful tool in determining potential instability, and it generally permits men and
equipment to be removed from harm’s way safely. The recent development of cost-
effective and sensitive instruments has made this process more accurate, repeatable
and reliable. For effective monitoring, a movement of a cm or less should be detected
to give a maximum time to react. Monitoring methods that can be used include:

• LiDAR Scanning stands for Light Detection and Ranging and uses a laser meas-
uring the reflection off surfaces to determine their precise position but relative
and not global. The accuracy is currently typically 1–3 cm.
• Global Positioning Systems use satellites and radio navigation techniques to
locate surfaces, etc. using at least four “line of sight” satellites to locate the sur-
face or point. The accuracy is typically in the range of 1m so has limited value for
accurate openpit stability monitoring currently.
• Slope Stability Radar Scanning uses radio waves to determine the position of
objects or surfaces. The most modern systems have accuracies in a mm or less
(e.g., the IBIS System).
• Time Domain Reflectometry is a measurement technique used to determine the
characteristics and disruption along electrical lines by observing reflected wave-
forms. It can be used with coaxial cables to determine failure and shear planes. It
is not a very cheap or versatile method as it can only determine movement across
a local cable often in a borehole. It is, however, able to determine very small
movements.

1.3.3 Rehabilitation/reclamation
After mining finishes, the mine area must undergo rehabilitation. Waste dumps are
contoured to flatten them out, to further stabilize them for the long term.
Reclamation plans include many of the following considerations:

• Adequate drainage control


• The preservation of the reclaimed topsoil layer
• Keeping waste segregated during mining
• Erosion and sediment control during and after mining
• Dust control
• Regrading to topography approved in the permit usually based on Environmental
Impact Study
• Restoration of mined areas

The plan must also consider the effects of any mine subsidence, vibration (induced
by mining, processing, transport or subsidence) and the impact on surface water and
groundwater. These environmental considerations often dictate the economics of a
planned mining operation and determine its viability.
If the ore contains sulphides, after mining, it is usually covered with a layer of clay
to prevent access of rain and oxygen from the air, which can oxidize the sulphides
to produce sulfuric acid, a phenomenon known as acid mine drainage. This is then
Surface mining 39

generally covered with soil, and vegetation is planted to help stabilize and consolidate
the material. Eventually this layer may erode over time, but it is generally hoped that
the rate of leaching or acid will be slowed by the cover such that the environment can
handle the load of acid and associated heavy metals. There are few long-term studies
on the success of these covers due to the relatively short time in which large scale
openpit mining has existed and been effectively closed. As mentioned earlier, it could
take decades or much longer for some waste dumps to become “acid neutral” and stop
leaching to the environment.

1.4 Geotechnical design


The geotechnical design consists of the following stages:

• Preliminary design
• Detailed design
• Design approvals
• Design implementation
• Design checking and modification based on the reality found.

Geotechnical design involves using soil and rock mechanics principles to design a
safe and economic mine design. Most designs need a reliable estimate of the rockmass
strength, which is governed mainly by local geology and hydrology.

1.4.1 Typical preparation and layout


The typical layout of an openpit involves both technical and economic considerations
and needs to be carefully investigated. The steps include:

• Purchase of all the mineral and related rights


• Exploration and geological drilling and other geophysical testing
• All permits and licenses obtained to start a mine
• Process plant design
• Block Model creation
• Economic evaluation, including the target production
• Final pit limit determination (which can change depending on the specific ore
value over time and any significant changes in the anticipated rockmass properties)
• Pushback (nested pit) designs. An openpit is mined in a series of pushbacks until
the ultimate pit limit is reached
• Pushback roadway designs
• Equipment selection
• Waste storage design
• Permit and environmental considerations
• Local community relations (that should start at the time of the mineral rights
purchase and continue until the rehabilitation is completed)
• Design optimization
• Continuous design modification based on the in-situ reality found with each push-
back and monitoring
40 Surface mining

Continuous design review is critical and this must also include the block grades as
well as the specific commodity price variation and predicted trend.
Fourie and Dohm (1992) outlined the entire openpit mine design process from
exploration to final design in Figure 1.34.
The stability of highwall benches in rocks is not generally a serious problem unless
major steeply dipping joints are present, in which case the highwall can be created at the
angle of the major joint set. If the rock has fine material inter-mixed (such as soil or clay),
then the slope or pile will behave more like a soil requiring the more conservative design.

Figure 1.34 The openpit mine design process (Fourie & Dohm, 1992)
Surface mining 41

1.4.1.1 Exploration
This is the most important phase and consists of geological drilling and other activities
such as geophysical testing. During the exploration phase, the following are established:

• The overburden width and stratigraphy


• The size, shape and boundaries of the orebody and any extensions
• The grade and distribution within the orebody
• The regional and local geology
• The hydrology
• The rockmass properties of both the orebody and host rock

From the detailed exploration data, a feasibility study (bankable document) is pre-
pared and the orebody is defined depending on the certainty, as follows:

• Inferred Reserves – Evidence suggests that there are minerals worth investigating;
sometimes described as “Potential Reserves”.
• Resource
• Indicated Resource: where initial drilling has identified that there is mineral-
ization but the precise configuration is uncertain.
• Measured Resource: where the tonnage has been calculated but drilling is not
sufficient to be certain of the orebody’s continuity and grade.
• Reserve – These are divided as follows:

• Probable Reserve: further testing has raised the level of confidence such that
obtaining initial funding can commence.
• Proven Reserve: where the orebody is well understood, and the tonnages and
grades established beyond reasonable doubt.

1.4.1.2 The final pit design


The pushback and final pit designs need to consider the chance of a failure (including
the size) and the consequence (such as could it adversely affect a main roadway or an
active mining area).
The final pit design involves the following considerations:

• The properties of the strata layers


• The hydrology within each layer
• The location and orientation of faults, major joints and any bedding
• Any surface infrastructure
• The climate
• Financial considerations

1.4.2 Geotechnical design processes


The geotechnical design of a surface mine involves both soil and rock mechanics
aspects over the entire life of the mining operation and the subsequent closure opera-
tions, to make it safe post-mining.
42 Surface mining

During the mine design process, the available data is limited and consists mainly of:

• Available literature from the respective government department or any previous


geological exploration
• Relevant outcrops
• Geophysical data
• Drilling core data
• Any test trenches or similar (often needed to finalize the processing plant design)

Often additional diamond drilling is required to further identify potentially adverse


geological or water features. These can influence the overall pit angle and the bench
height and bench angle.
The Department of Minerals and Energy (DME) for Western Australia (1999) rec-
ommended that the following must be considered in their document titled Geotechnical
Considerations in Openpit Mines: Guidelines:

• Depth and operating life of the mine


• Any potential for changes in expected ground conditions in the wall rockmass
(e.g., rock strength, earthquake events, rock stress, rock type, etc.)
• The production rate
• Size, shape and orientation of the excavations
• The location of major working benches and transportation routes
• Potential for surface water and groundwater problems
• The equipment to be used, excavation methods and handling of ore and waste
• The presence of nearby surface features (e.g., public roads, railways, pipelines,
natural drainage channels or public buildings)
• The potential for the general public to inadvertently gain access to the mine void
during and after mining
• Time-dependent characteristics of the rockmass (particularly after abandonment)

Hydrology plays an important role in the stability of an openpit and water is always
the “enemy” in mining. Hoek and Bray (1981) listed the adverse effects of water in an
openpit to be:
• An increase in pore pressure within the rockmass (which reduces shear strength)
The effect of pore water is significant on the reduction of shear strength as indi-
cated in the Mohr-Coulomb equation below:
τ  = So − µ( σ − Po )

Where: τϴ is the shear strength, So is cohesion, μ is the angle of internal friction, σϴ


is the stress and Po is the pore water pressure.

• Softening of infill or rock material


• Erosion of weaker bands of rock by water seepage or egress
• Weathering or slaking of soft rock due to wetting and drying cycles
• Reduced blasting efficiency
• Corrosion of steel support and reinforcement
Surface mining 43

The DME (1999) recommended the following geotechnical site-specific approach,


again based on their publication titled Geotechnical Considerations in Openpit Mines:
Guidelines:
• Use of scanline or other methods of geotechnical wall mapping and/or oriented core
logging to establish baseline geotechnical data on planes of weakness within the
rockmass with a minimum of bias. This approach is particularly useful when mines
are developed as a series of cutbacks. Wall mapping should attempt to quantify
orientation, persistence, spacing, roughness and wavelength, wall rock strength,
aperture, infill, degree of weathering and moisture content of planes of weakness.
• Take representative samples of the rockmass and determine relevant shear
strengths and groundwater characteristics.
• Use of (preferably computer-based) plotting, analysis and presentation methods
of geological structure data to better define orientation, persistence, spacing and
other characteristics of joint sets.
• Identify the general geotechnical domains in the rockmass throughout the mine.
• Transfer of this data to geological plans and/or computer models for use in the
design of pit walls and wall support and/or reinforcement (where required).
The shear strength of the rockmass and geological features are important when designing
an openpit mine.
The key to safe and cost-effective exploitation of the orebody is to continually revisit
the design based on the ever-increasing geotechnical database. One of the most impor-
tant functions is to monitor pit wall movement and mapping of the faces considering
the orientations of the faults and joints and their condition. The effects of blasting
must also be evaluated for the effectiveness of the muckpile position and fragmenta-
tion, and any resultant damage to the benches.
Blast design must be done on a site-specific basis and optimized with time as more
experience is gained. The blasting may need to be modified based on the different
rockmass properties around the pit. Most suppliers of explosives and blasting ancil-
lary products offer a design and monitoring service that can be useful.
According to the DME (1999), in their publication titled Geotechnical Considerations in
Openpit Mines: Guidelines.
The factors that control the level of highwall damage caused by drilling and blasting
include:
• Rockmass properties such as orientation, persistence and spacing of geological
structures and the presence of groundwater (i.e., rockmass classification)
• The degree of “confinement” and amount of burden shifted by the proposed blast
• Inadequate removal of rock debris from earlier blasts from the toe of batter slopes
(sections of rockmass between catch berms within pit walls – usually excavated to
a specific inclination/angle from the horizontal)
• The degree of rock fragmentation required
• Selection of the appropriate hole diameter
• Control of individual hole collar position, hole bearing, inclination and length
• The type and amount of stemming used
• Placement of holes in a suitable pattern to achieve the required excavation geometry.
• The use of specific perimeter holes such as stab holes, or smooth blasting tech-
niques (e.g., pre-splitting, post-splitting or cushion blasting)
• Selection of appropriate initiation system(s) and initiation sequence of the blast or blasts
44 Surface mining

• Specific types or combinations of explosives. Explosives must be selected accord-


ing to the given ground mass conditions, e.g., groundwater or reactive shales can
affect the result of a blast. Explosives must also be selected to achieve the required
energy levels, maintain compatibility with the initiation systems and the explo-
sives’ expected useful life in blast holes (which can be reduced mainly by water)
• Control of explosive energy levels in the near-wall holes and preferably using
decoupled explosive charges, with a cartridge diameter less than the blast hole to
minimize blast damage at the excavation perimeter
• The required mining bench height and the depth of subgrade drilling (subdrill)
• Availability of well-maintained drilling, explosives handling and charging equip-
ment of appropriate capacity
A slope design methodology (Haines et al, 2006) is given in Figure 1.35.

Figure 1.35 Slope design methodology (Haines et al., 2006)


Surface mining 45

Support is generally avoided if possible and based rather on the stable “unsup-
ported” design, but localized support can be installed where conditions require it,
especially up-dip from major roads. Support also becomes more of a consideration in
seismically active areas.
Hoek and Marinos (2000) show the influence of jointed rock on rockmass strength
using the Geological Strength Index (GSI) in Figure 1.36.

Figure 1.36 The effect of GSI (Hoek and Marinos 2000)


46 Surface mining

1.4.3 Geotechnical models


Geotechnical models are usually based on rockmass classification that can determine
the geotechnical domains. The following are the most commonly used systems:

• Rock Quality Designation (RQD), which is a fundamental part of the more


detailed and representative methods (Deere, 1968)
• Rockmass Rating system (RMR) (Bieniawski 1976)
• Mining Rockmass Rating system (MRMR) (Jakubec & Laubscher 2000)

There are many design methods that can be used, but the essential requirement is hav-
ing reliable and representative input parameters (otherwise “rubbish in equals rubbish
out”). The design methods that can be used include:

• Experiential methods
• Empirical methods
• Mathematical methods
• Computer/numerical methods

1.4.3.1 Experiential design


Experiential design is based on the person’s first-hand relevant prior experience. It is
useful to help focus on the real site-specific issues and can be used to determine a “first
brush” rough design or to intuitively check another design. It cannot be relied upon
for a final design.

1.4.3.2 Empirical design


The word “empirical” is derived from the Greek word for “experience” (εμπειρία).
Empirical design uses a database of similar situations and establishes a cut-off for
success and unsuccessful designs within certain parameters. The basis of empirical
design is previous relevant experience, without consideration for any systematic the-
ory. Most pillar designs used in mining are empirically based for example.
The empirical design techniques for an openpit mine are generally involved with
determining the rockmass characteristics and identifying all the geotechnical domains
that will be encountered during the openpit life.

1.4.3.3 Mathematical design


Mathematical design can be used for individual localized potential falls or slides. The
mathematical design is therefore not applicable for holistic pit designs.
Terzaghi (1943) divided the causes of slope failures into “internal” and “external”
causes.

• Internal causes – the failure occurs without external change (including earth-
quakes) and is usually a result of increased pore water pressure.
• External causes – the shearing stress is increased without the shearing resistance
of the material changing. This shear stress increase can typically be caused by
steepening the slope angle, increasing the slope height or by erosion.
Surface mining 47

Slope stability analysis is based on assumptions that were developed over a 100 years
ago and include:

• Slope failure occurs along a particular sliding surface/plane and thus the failure
can be assumed to be a two-dimensional plane problem.
• The failure mass moves as a rigid body and any deformations (if they occur) do
not influence the analysis.
• The shearing resistance of the soil mass at points along the failure surface is
independent of the orientation of the failure surface (strength properties are
isotropic).
• The factor of safety is defined in terms of the average shear stress along the poten-
tial failure plane and the average shear strength along the plane, rather than spe-
cific values at specific locations.

A typical slope stability geometry is given in Figure 1.37


In slopes, angle (β) and slope vertical height (H) (as shown in Figure 1.37) can be of
interest for the following reasons:

• The steepest stable angle for any given height


• For a given slope angle and height, the factor of safety

The following gives examples where a mathematical solution can be used to model
stability in rock or soil localized.

• A cohesive slope material

A purely cohesive material is one where the cohesive strength (s) is not zero and the
angle of internal friction (φ) is zero (see Figure 1.38).
Creq is the cohesive (resistive) force on the potential failure surface required to
maintain equilibrium (i.e., stop sliding failure). Based on statics:

C req = W * Sin i (1)

Figure 1.37 A typical slope with the main design parameters


48 Surface mining

Figure 1.38 Slope with a potential failure plane

The weight of the failure block W is:

W = 12 * γ * (H/Sin i) * (H/Sin β ) * [Sin (β − i)] (2)

Where:

γ is the material weight density


(H/Sin i) is the length ac
(H/Sin β) is the length ab
[Sin (β – i)] is (bb’/ab) from Figure 1.38.

If “s” is the average shearing stress along the failure surface, then:

Creq = s * (H/Sin i) (3)

By substitution of equations (2) and (3) into equation (1), the following is obtained:

s(H/Sin i) = 1
2 * [( γ * H2 )/(Sin i * Sin β )] * [Sin(β − i)] * Sin i

Which can be simplified to:

s/( γ * H) = [Sin i * Sin(β − i)]/2 * Sin β (4)

s/(γ * H) is called the stability number and is dimensionless. It is the ratio of the cohe-
sive component of shearing resistance to γH that is required to maintain stability (i.e.,
factor of safety of 1).
From a slope stability point of view therefore, the most critical plane is that where
(s/γH) is maximum.
Differentiating with respect to angle i gives:

∂ / ∂ i(s/γ H) = ∂ / ∂ i[Sin i * Sin(β − i)] = 0 from which: i = 1


4 * Tan(β /2)
Surface mining 49

Therefore:

s/(γ * H) = 1
4 * Tan(β /2) (5)

Where:

s is in kg/m2, γ is in kg/m³ and H is in m

Equation (5) can therefore be used to find the maximum slope H with a given “s” and
slope angle β or a maximum angle β for a given “s” and H.

• A soil-based slope material

Soil slope failure is well known but still requires some trial and error and several iter-
ations to identify the potential failure arc. The failure surface of a typical flow failure
can be approximated as a rotational collapse of a circular arc as shown in Figure 1.39
(Taylor, 1948).
To analyze a planned or existing slope before failure, a worst (largest) failure surface
is assumed, as shown in Figure 1.40.
The failure surface as shown in Figure 1.40 is also assumed to be in a uniform soil.
Generally, the failure mass is assumed to have unit thickness.

Figure 1.39 Rotational slope failure (Taylor, 1948)


50 Surface mining

Figure 1.40 Worst case rotational failure slope (Taylor, 1948)

The factor of safety:


SF = (R * s * L′ * b)/(γ * A * b * d) = (R * s * L′ * b)/(W * d)

Where:
R is the radius of the failure plane
s is the average unit shearing strength
Ĺ is the arc length of the sliding/failure surface
b is the unit thickness (generally unity)
γ is the soil unit weight
A is the sliding mass area
d is the moment arm of the sliding mass
W is the weight of the sliding mass with thickness/depth of b

This type of analysis only establishes the factor of safety for that specific assumed
failure surfaces, and in practice, there are many such potential failure surfaces. The
one of interest is the failure surface that gives the lowest factor of safety. The failure
surface with the lowest factor of safety is found usually by trial and error.
A typical example is shown in Figure 1.41.
Assume that a failure surface (A) was first assumed and the factor of safety was
found to be 1.24. Another failure surface (B) to the right of (A) is now selected and
the factor of safety is found to be lower at 1.14. A third failure surface (C) is selected
even further to the right and the factor of safety is found to have increased to 1.27. The

Figure 1.41 Different rotational failure surfaces (Taylor, 1948)


Surface mining 51

Figure 1.42 The potential failure surfaces based on the slope angle (Taylor, 1948)

failure surface (B) is therefore assumed to be potential failure surface and the slope is
assumed to have a factor of safety of 1.14 and therefore stable.
The method described below can therefore be used to find the potential failure surface
by this “trial and error” method. The method is obviously highly simplified, and apart
from the slope material often being variable, water is the most critical consideration.
The potential failure surface also changes with the slope dip as shown in Figure 1.42.

• Flat slopes of 3:1 horizontal:vertical (18.5°) have a worst circle/surface with high
curvature dipping down into the slope base.
• Moderate slopes of 2:1 to 1.5:1 ratio (21.8°–34.7°) have a worst circle with less cur-
vature passing through the slope toe.
• Steep slopes (+34.7°) the worst circle is flat and can be approximated as a plane surface.
• Slopes of broken rock.

Broken rock (spoil) will naturally stand at the angle of repose of the material. An
important factor in rock piles is that the stability is independent of the height of the
rock pile. The typical angle of repose varies from 35° to 40°.

1.4.3.4 Computer design


Computer design has become the most effective as long as care is taken with the input
parameters. Sensitivity analyses should also be undertaken, which is very useful.
Many different commercially available packages are available; the key is validating
the model. Once mining starts, this can be done using field measurements so the model
becomes very effective and quantitatively close to reality.
52 Surface mining

Depending on the site-specific complexity, two- or three-dimensional modelling


can be undertaken on each geotechnical domain around the pushbacks and to the
final pit boundary.
Commonly used software for rock mechanics and/or scheduling includes:

• DIPS (RocScience)
• SLIDE (RocScience)
• UDEC or 3DEC (ITASCA Consulting Group)
• FLAC or FLAC3D (ITASCA Consulting Group)
• Micromine
• Vulcan (Maptek)
• Surpac
• Datamine
• Geomine Quickpit (Threedify)
• Minesight
• Whittle

1.4.4 Geotechnical bench and slope design analysis techniques


The overall strength of a rockmass is mainly governed by the strength, continuity and
properties of the fractures, cracks or joints. The strength of the intact rock can be also
important particularly if it tends to be relatively weak or prone to weathering.
There are many algorithms that can be used when developing the pushbacks, and
one of the first was by Lerchs and Grossmann (1965). The goal is always the same –
maximizing the NPV of the orebody safely.
The bench is designed taking into account the bench height, width (equipment
related) and angle (Figure 1.43). Often a catch bench is left during a pushback to con-
trol unplanned fallen rock and facilitate access as shown in Figure 1.44.
As mentioned, the bench height depends mainly on the rockmass, geology and the
equipment used. The higher the bench, the lesser the selectivity. Mohr-Coulomb equation

Figure 1.43 Bench design and definitions (Call, 1986)


Surface mining 53

Figure 1.44 A catch bench design (Call, 1986)

is often used to determine a safe bench height from a rock mechanics perspective, and
the prominent joint sets are often the deciding factor of the bench angle. Generally the
steeper, the better economically, but safety is still the overriding consideration.
A factor of safety of 1.2–1.4 is generally used (Willie and Mah, 2005).
The geotechnical design process is shown in Figure 1.45 after Trongvann, 2019.
The condition of the slope base (floor) is also important, and the presence of weak
material (clay) and water can cause serious foundation problems, which can affect the
design and stability.

1.4.5 Slope monitoring


Slope monitoring should be undertaken visually and using real-time monitoring
systems.
Visual observation can identify:

• The widening of tensile cracks and fractures


• Water seepage
• Highwall bulging or bowing
• Unravelling of the rock
• Bending or failure of rock supports

Slope monitoring can be done by:

• Remote instrumentation (using satellites, radar, lasers, etc.)


• Photogrammetry
• Survey prisms
• Extensometers (site specific)
• Borehole inclinometers or tiltmeters (site specific)
• Displacement monitors (site specific)
54 Surface mining

Figure 1.45 The geotechnical design process (Trongvann, 2019)

The results of a failure can be catastrophic so a comprehensive and reliable moni-


toring program is absolutely essential. Losses can include fatalities and injuries, dam-
age to equipment and/or loss of production.
The terms and definitions are largely based on:

• ISRM Glossary of terms isrm.net/isrm/glossary/index?msclkid=9a1dd59ab


99911ecaf508b24d33d255f
• Rockburst Terminology – South African Chamber of Mines Circular 28/82.
• RPM Global’s Glossary of Mining Terms.

References
Arteaga, F., Nehring, M., Knights, P. & Camus, J. (2014) Schemes of Exploitation in Open Pit
Mining. Mine Planning and Equipment Selection Conference (January).
Bieniawski, Z.T. (1976) Rock mass classification in rock engineering. Proceedings of the
Symposium on Exploration for Rock Engineering, Cape Town, Balkema.
Calder et al. (1997) Planificacion Operativa de Minas a Cielo Abierto – International Specialization
Program in Mining Technology. Pontificia Universidad Catolica de Chile (unpublished material).
Call, R.D. (1986) Cost Benefit Design of Openpit Slopes. Call & Nicholas Publications.
Surface mining 55

Camus, J. (2002) Management of Mineral Resources: Creating Value in the Mining Business.
Society for Mining, Metallurgic and Exploration. ISBN: 0-87335-216-5.
Dagdelen, K. (2000) Open Pit Optimization: Strategies for Improving Economics of Mining
Projects through Mine Planning. Application Computers for Mining Industry.
Deere, D.U. (1968) Geological considerations. In: Stagg, K.G. & Zienkiewicz, O.C. (eds) Rock
Mechanics in Engineering Practice. John Wiley and Sons, New York, pp. 1–20.
Dyno Westfarmers Ltd. (1993) Efficient Blasting Techniques. Sanctuary Cove.
Fourie, G.A. & Dohm, G.C. (1992) Mining Engineering Handbook, Open Pit Planning and Design
Chapter 13.1, 2nd Edition, pp. 1274–1290.
Hagan, T.N. & Mercer, J.K. (1983) Safe and Efficient Blasting in Open Pit Mines. Manual written
for the course given at Karratha, Australia (November) ICI Australia Operation Ltd.
Haines, A., Voulgaris, P., Walker, D. & de Bruyn, I. (2006) Geotechnical design considerations for
the proposed you Tolgoi openpits, Southern Mongolia. International Symposium on Stability of
Rock Slopes in Open Pit Mining and Civil Engineering, SAIMM, Johannesburg, South Africa.
Hartman, H.L. (1992) SME Mining Engineering Handbook. Society for Mining Metallurgy &
Exploration.
Hartman, H.L. & Mutmansky, J.M. (2002) Introductory Mining Engineering. John Wiley & Sons.
Hoek, E. and Bray, J.W. (1981) Rock Slope Engineering. Stephen Austin and Sons Limited, Hertford.
Hoek, E. & Marinos, P. (2000) Predicting tunnel squeezing problems in weak heterogeneous rock
masses. Tunn Int, Vol. 32, No. 11, pp. 45–51.
Jakubec, J. & Laubscher, D.H. (2000) The MRMR Rock Mass Rating Classification System in
Mining Practice. MASSMIN 2000. Brisbane, Australia. November.
Lerchs, H. & Grossmann, I.F. (1965) Optimum Design of Open Pit Mines. Joint CORS and
ORSA Conference. Transactions, CIM. Montreal, Canada. pp. 17–24.
Mines Occupational Safety and Health Advisory Board (1999). Department of Minerals &
Energy, Perth, Western Australia.
Prajapati, G. (2016) Soil Mechanics: Types of Slope Failures. Indian Institute of Technology.
Schissler, A.P. (2004) Coal mining, design and methods. Encyclopedia of Energy. pp. 485–494.
Somagani, P.C. (2016) Stochastic pit optimization under slope stability uncertainty. Final year
thesis. University of Western Australia. DOI:10.13140/RG.2.2.33027.81444 - May.
Songolo, M. (2010) Pushback design using genetic algorithms. Masters thesis. Western Australia
School of Mines. DOI:10.13140/2.1.4065.6169 - June.
Tannant, D. & Regensburg, B. (2001) Guidelines for Road Haul Mine Design. UBC Open Collections.
Taylor, D. (1948) Fundamentals of Soil Mechanics. John Wiley and Sons.
Terzaghi, K. (1943) Theoretical Soil Mechanics. John Wiley and Sons, New York. ISBN
0-471-85305-4.
Trongvann, N. (2019) Blog-Geotechnical Considerations for Openpit Design. Earth & You. December 21.
Warner, R.K. (1934) Mining Systems-Selection of a Mining System. The American Institute of
Mining, Metallurgical, and Petroleum Engineers. January.
Westcott, P., Pitkin, G. & Aspinall, T. (2009) Open-cut mining. In: Kininmonth, R.J. & Baafi, E.Y.
(eds) Australasian Coal Mining Practice, Chapter 18, Monograph 12, 3rd Edition. AusIMM,
Melbourne, Victoria, Australia, pp. 410–458.
Whittle, J. (1989) The Facts and Fallacies of Open Pit Optimization. Whittle Programming Pty
Ltd., Australia.
Willie, D.C. & Mah, C.W. (2005) Rock Slope Engineering. Taylor and Francis, New York, p. 456.
Chapter 2

Underground soft rock mining

2.1 Introduction to soft rock mining

2.1.1 Coal formation and geology


Coal is an organic material that forms in sediments from large volumes of vegetation
as they die, regrow (frequently in swamps and wetlands) and are slowly buried over
geological time becoming solidified due to increasing temperature and pressure. The
swamp location is important as the dead vegetation does not always readily decay as
the swamp is typically oxygen deficient, so peat is often formed and then converted to
coal over several hundreds of millions of years. The simplified Figure 2.1 shows the
formation of peat and then the various coal grades.
Coal consists typically of:

• Carbon
• Water
• Silicate minerals
• Sulphur compounds
• Clay materials

Depending on the coal grade, it is typically used for:

• Electricity generation (the largest use)


• Heating
• Steel formation
• Gas, tar and oil formation typically using distillation
• Fertilizer production
• Jet fuel

Many countries have mineable coal reserves, and it was the most cost-effective elec-
tricity producer in the last century. However, global production is dropping due
mainly to climate change considerations and alternatives that have become more
cost effective.
Table 2.1 shows the major coal reserves by country.
At the present annual coal consumption (which is likely to reduce over time), the
world has over 130 years of coal reserves.

DOI: 10.1201/9781003185680-2
Underground soft rock mining 57

Figure 2.1 The formation of coal (Surawar – 2014)

Table 2.1 The global coal reserves by country

Country Coal reserve (million t) Global reserve share Region

US 249,000 23% North America


Russia 162,000 15% Asia/Europe
Australia 150,000 14% Oceania
China 143,000 13% Asia
India 111,000 10% Asia
Germany 36,000 3% Europe
Indonesia 35,000 3% Asia
Ukraine 34,000 3% Europe
Poland 28,000 2% Europe
Kazakhstan 26,000 2% Asia
Turkey 12,000 1% Europe
Rest of the world ≈120,000 11%

Source: BP Statistical Review of World Energy (2021)


58 Underground soft rock mining

2.1.2 Coal classification


Coal is classified into four main types:

• Peat (basically “pre-coal”)


• Lignite
• Sub-bituminous coal
• Bituminous coal
• Anthracite

The coal type depends on the amount of carbon it contains and on the amount of heat
energy it can produce when burnt. This depends on the pressure and heat acting on the
plant debris as it sank deeper and deeper over millions of years (as shown in Figure 2.2).
The age of coal deposits differs depending on the global location as shown in Figure 2.2.
Coal usage varies depending on the grade as shown in Figure 2.3.

Figure 2.2 Coal age based on global location (Energy Resources: Coal, Open Universities)

Figure 2.3 Coal usage depending on its grade


Source: Dept of Industry, Innovation and Science, Australia
Underground soft rock mining 59

2.1.2.1 Peat
Peat is a brown, soil-like material characteristic of boggy, acid ground, consisting
mainly of decomposing partially carbonized vegetation (see Figure 2.4). It typically
has a carbon content of less than 60% based on a dry ash free analysis.

Figure 2.4 Peat samples


Photograph (a) reproduced with kind permission from the Mining Museum at China University of Mining &
Technology.
Photograph (b) reproduced with kind permission from K. Barmore. Object [5402] W. M. Keck Earth Science
and Mineral Engineering Museum University of Nevada, Reno.
60 Underground soft rock mining

Figure 2.5 Lignite sample


Photograph reproduced with kind permission from K. Barmore. Object [7256] W. M. Keck Earth Science and
Mineral Engineering Museum University of Nevada, Reno.

Major peat-producing countries are Canada, Germany, Latvia and Ireland. The
Netherlands used to also be a significant producer (Barasa, 2020).

2.1.2.2 Lignite
This is the lowest rank of coal and with the lowest energy content. Lignite coal depos-
its tend to be relatively young coal deposits that have not been subjected to extreme
heat or pressure. Lignite is a weak and crumbly rock with a relatively high moisture
content as shown on Figure 2.5. It typically has a carbon content of 60–70% based on
a dry ash free analysis. Lignite is mainly burned at power plants to generate electricity.
Garside (2020) estimated global lignite production to be about 800 million metric
tonnes in 2018. Countries with significant lignite reserves include Russia, Australia,
Germany, US, Indonesia, Turkey, China, New Zealand and Poland.

2.1.2.3 Sub-bituminous and bituminous coal


Sub-bituminous coal: Sub-bituminous coal is formed from lignite that has been subjected
to further metamorphism (from increased temperature and pressure) which has reduced
the hydrogen and oxygen content. It therefore has a higher heating value than lignite.
Sub-bituminous coal typically contains 70–78% carbon based on a dry ash free analysis.
It is mainly used to generate electricity in thermal power plants. The Powder River Basin
in Wyoming, US has massive sub-bituminous coal seams (heights of ~20m) that out-crop
at or near the surface.
Underground soft rock mining 61

Figure 2.6 Bituminous coal


Reproduced with kind permission from the Mining Museum, China University of Mining & Technology

Most sub-bituminous coal in the US is at least 100 million years old, and about 42%
of the coal produced in the United States is sub-bituminous.
Bituminous coal: Bituminous coal was formed under even higher heat and pres-
sure. It has two to three times the heating value of lignite. Bituminous coal is used to
generate electricity and is an important fuel and raw material for the steel and iron
industries. Bituminous coal typically contains 77–87% carbon based on a dry ash free
analysis (Figure 2.6).
Bituminous coal is the most common grade of coal produced globally, and annual
production is shown in Figure 2.7.

2.1.2.4 Anthracite
Anthracite is much harder and has a higher density than bituminous coal, and unlike
other coal grades, it is considered a metamorphic rock (as shown in Figure 2.8).
Anthracite typically contains over 87% carbon based on a dry ash free analysis and
62 Underground soft rock mining

Figure 2.7 Coal historical and future production


Source: Li, M. (ed.), Coyne, D. (2018) World coal – 2010–2050 World Energy Annual Report, Part 4 –
September

Figure 2.8 Anthracite sample


Photo reproduced with kind permission from the Mining Museum, China University of Mining & Technology
Underground soft rock mining 63

Figure 2.9 The global anthracite market and forecast

a low percentage of volatile material. It has a heating value slightly lower than bitu-
minous coal. It is considered an intermediate product between bituminous coal and
graphite.
Anthracite accounts for about 1% of total coal reserves (Kay, 2018). China is the
largest producer and other producers include Russia, Ukraine, North Korea, South
Africa, Vietnam, the UK, Australia, Canada and the US. The global anthracite mar-
ket is shown on Figure 2.9 (Maximize Market Research, 2020). It is, however, very rare
in the US, accounting for less than 0.5% of the coal mined in the US. All the anthra-
cite mines in the US are located in northeastern Pennsylvania and tend to be in very
narrow seams.
World coal reserves are given in Table 2.2.

2.1.3 Coal exploration


Before coal exploration begins, the exploration licenses, titles and/or options must be
obtained.

Table 2.2 World coal reserves (Gaedicke et al., 2019)

Anthracite and Lignite and Continent


Countries with high bituminous coal sub-bituminous coal Total
Continent/Region coal reserves (1,000 t) (1,000 t) (1,000 t)

North America US 220,167 30,052 258,012


South America 14,016
Europe Poland 20,542 5,937
Ukraine 32,039 2,336 135,593
CIS Kazakhstan 25,605 0
Russia 68,634 90,730 188,853
Africa and Middle South Africa 9,893 0 14,420
East
Asia Pacific Australia 70,927 76,508
China 130,851 7,968
India 96,468 4,895
Indonesia 26,122 10,878 444,888
Global Total 1,054,782
64 Underground soft rock mining

The aim of coal exploration is to gather enough information about the deposit
so a decision can be made on the viability of the deposit for safe and profitable
mining. The quantity and quality of the coal deposit, as well as the potential ease
of mining it, which is mainly influenced by the geology and hydrology must also be
determined.
In the exploration phase, the goal is to determine whether the deposit can be turned
into a viable mine and provides the information needed for:

• The Environmental Impact Study.


• Determining the mining method, mainly between surface or underground min-
ing and if underground, whether room and pillar or longwall. This also requires
determining the market for the product and the costs and net present value of the
proposed new mine.

Typical exploration activities include:

• Literature review
• Geological mapping
• Photos from drones, aircraft and satellites
• Geophysical surveys (gravity, magnetic and seismic surveys)
• Diamond drilling
• Initial contact with the local community

Diamond drilling is the costliest exploration activity but provides the most detailed
information, such as:

• The stratigraphy
• The rock properties of the overburden, the coal seam and the floor
• The regional and local geology and hydrology
• The coal chemistry, thickness and calorific properties

2.1.4 Typical soft rock mining


Soft rock mining includes coal, halite, trona and sulphur and generally requires the
minimum of drilling and blasting, unless it is a surface operation.
Typical mining methods include:

• Strip mining
• Solution mining
• Highwall mining
• Room and pillar
• Longwall mining

Underground soft rock mining frequently uses continuous miners and road headers
rather than drilling and blasting.
Underground soft rock mining 65

Figure 2.10 A continuous miner


Reproduced with kind permission from Komatsu Mining Corp.

Figure 2.10 shows a continuous miner (note the gathering arms to collect the broken
coal and transfer it onto the short chain conveyor to load onto typically a shuttle car
or ramcar).
Figures 2.11 and 2.12 show a road header, which has an advantage over a continuous
miner because it can cut any excavation profile whereas a continuous miner can only
cut a square or rectangular profile, but is much more productive.

Figure 2.11 A road header


Reproduced with kind permission from Antraquip
66 Underground soft rock mining

Figure 2.12 A road header cutting a tunnel


Reproduced with kind permission from Antraquip

2.2 Basic coal (soft rock) mine design and planning

2.2.1 Coal mine planning and design principles


According to Fourie and van Niekerk (2001), the goal of planning and design of any
coal mining operation is to design an integrated coal exploitation system that will
ensure that the coal is extracted and prepared to meet a desired market specification,
at a minimum unit cost, within acknowledged safety, health, social, legal and regu-
latory constraints. A large number of specific engineering, scientific and economic
disciplines contribute interactively to the overall mine planning and design process,
thus making it a true multi-disciplinary activity.
They suggest that any mine planning and design process typically consists of the
following five unique and identifiable phases:

• Phase 1: Project data collection and investigations


• Phase 2: Evaluation, planning and design
• Phase 3: Construction and mine establishment
Underground soft rock mining 67

• Phase 4: Mining operations


• Phase 5: Mine decommissioning and closure

Each of the above phases, they state, should consist of the following activities (modi-
fied by the authors):

• Project Phase desired outcome definition


• Statement of all assumptions used for all of the planning phases
• Identification of all planning and design risks
• Identification of planning and design restrictions and constraints from any source
(regulatory, environmental, community, funding, infrastructural, climate, etc.)
• All planning and design criteria to be used and the justification
• Relevant data collection
• System planning
• Hazard identification, risk assessment and management
• Preparation of various options
• Option evaluation
• Identification of the best or preferred options based on safety, production and
profitability
• System design
• Ongoing review and updating of the design as new information is available

2.2.2 Coal mine rockmass classification


The Coal Mine Roof Rating (CMRR) is an example of a coal rockmass classifica-
tion method and was developed by Mark and Molinda (2005) to meet the needs of
mine planners for a simple, repeatable and meaningful classification system of US
coal mines. It employs the familiar format of Bieniawki’s Rock Mass Rating (RMR),
including summing the individual ratings to obtain a final CMRR on a 0–100 scale. It
is also designed so that the CMRR unsupported span and stand-up time relationship
is roughly comparable to the one determined using the RMR. This section is based
directly on Mark and Molinda’s work in developing and validating the CMRR.

2.2.2.1 Data collection for calculation of CMRR


The data required for the CMRR can be determined either from observations of under-
ground exposures such as roof falls and overcasts, or from exploratory drill core (or both).
In either case, the main parameters required are:

• The uniaxial compressive strength (UCS) of the intact rock types


• The intensity (spacing and persistence) of bedding and other discontinuities
• The shear strength (cohesion and roughness) of bedding and other discontinuities
• The moisture sensitivity of the rocks
• The presence of a prominent bed in the bolted interval

Other, secondary, factors include the number of rock layers, the presence of ground-
water and surcharge from overlying weak beds.
68 Underground soft rock mining

2.2.2.2 The CMRR calculation process


The CMRR is calculated in a two-step process:

• In the first step, the mine roof is divided into structural units, and Unit Ratings
are determined for each. A structural unit generally contains one lithologic layer,
but several rock layers may be lumped together if their engineering properties are
similar.
• In the second step, the CMRR is determined by averaging all the unit ratings
within the bolted interval (with the contribution of each unit weighted by its thick-
ness) and applying appropriate adjustment factors. This second step is the same
regardless of whether the Unit Ratings were from data collected underground or
from core. Figure 2.13 outlines the process.

2.2.2.3 Parameters involved in the calculation of CMRR


The following parameters are required:

• UCS:

The UCS of the rock in the roof is important because:


• It determines the ease with which new fracturing (as opposed to movement
along pre-existing discontinuities) will take place.
• The compressive strength of the rock is a factor in the shear strength of
discontinuities.
The compressive strength rating contributes about a third to the CMRR. This is con-
siderably more than the weight it contributes to the original RMR.
Laboratory testing is generally considered the standard method of determining the
UCS. Unfortunately, laboratory tests of weaker rocks can be more costly and tedious
to prepare and test. The variability in the UCS also tends to be high, with the standard
deviation typically about one-third of the mean for coal measure rocks.

Figure 2.13 Steps involved in calculation of CMRR (Mark & Molinda, 2005)
Underground soft rock mining 69

Figure 2.14 CMRR rating scale for Axial Point Load or UCS tests

As an alternative for the CMRR, Clark and Molinda (2007) recommend the Point
Load Test (PLT) for drill core. An advantage of the PLT is that numerous tests can be
performed quickly and cheaply, because the procedures are simple. The test appara-
tus is also inexpensive and readily portable. Another advantage of the PLT is that
both diametral and axial tests can be performed on any core. In a diametral test,
the load is applied parallel to bedding. The diametral test is therefore an indirect
measure of the lateral strength, or bedding plane shear strength. When the axial
PLT is used to estimate the UCS, the Point Load Index (Is50) is converted using the
following equation:
UCS = K * ( Is50 ) (1)

Where K is the conversion factor.


A comprehensive study involving more than 10,000 tests of coal measure rocks
reported that a value of K = 21 fits the data for the entire range of rock types and geo-
graphic regions tested. The UCS rating scale used in the CMRR program is shown in
Figure 2.14 (Mark & Molinda, 2005).
Underground, the CMRR employs an indention test to estimate UCS. The exposed rock
face is struck with the round end of a ball peen hammer, and the resulting characteristic
impact reaction is used to calculate CMRR rating for UCS as given in the Table 2.3. The
nature of the reaction (depth of indentation), not its magnitude, is what is important.

Table 2.3 CMRR rating scale for ball peen hammer test

Ball peen hammer class CMRR rating

Molds 5
Craters 10
Dents 15
Pits 22
Rebound 30
70 Underground soft rock mining

• Discontinuity Intensity – For CMRR, Mark and Molinda (2005) state that the
discontinuity intensity is determined by the spacing between bedding planes or
other discontinuities, and the persistence, or extent, of each individual disconti-
nuity. The more closely spaced a set of discontinuities, the greater the weakening
effect it has on the rock mass. Persistence is more important for discontinuities
that are widely spaced. Like UCS, the Discontinuity Intensity accounts for about
one-third of the total CMRR.
Table 2.4 Bedding and discontinuity intensity table for underground data

Spacing

Persistence (m) >1.8 (m) 0.6–1.8 (m) 0.2–0.6 (m) 60–200 (mm) <60 (mm)

0–1 35 30 24 17 9
1–3 32 27 21 25 9
>3 30 25 20 13 9

Underground, both spacing and persistence can be measured directly, using the
standard methods for rockmass characterization. Table 2.4 shows the Bedding/
Discontinuity Rating Scale for underground data. The matrix shows what point
value is added for each combination of spacing and persistence of discontinuities.
Most standard geotechnical core logging procedures include considering some
measure of the natural breaks in the core. The two most commonly used methods
are the fracture spacing and the Rock Quality Designation (RQD). Fracture spac-
ing is easily determined by counting the core breaks in a particular unit, and then
dividing by the thickness of the specific unit. The RQD is obtained by dividing
combined length of core pieces that are greater than 100mm in length by the full
length of the core run (including the length of any core losses, but excluding any
fresh fractures caused by drilling).
• Shear Strength of Discontinuities – Bedding plane shear strength is a critical
parameter for coal mine underground stability because the most severe loading
in the coal mine roof is normally lateral, mainly caused by high horizontal stress,
especially in the US. The shear strength of discontinuities rating is evaluated using
two parameters: cohesion between the bedding surfaces and the roughness of the
surfaces. Underground, the cohesion of bedding surfaces is estimated by using
a 9-cm mason chisel and a hammer to split hand samples of rock. Cohesion can
also be estimated by observing the nature of the fractured rib (wall) of a roof fall.
The roughness along a discontinuity surface is the other component of the rock
surface’s shear strength. In CMRR, roughness of a surface is estimated visually and
roughly classified into “jagged”, “wavy” or “planar”. This measure is to be applied
on a scale which ranges from hand sample size to several metres across a fall expo-
sure. The CMRR assumes that roughness significantly affects shear strength only
when cohesion is in the middle range. Table 2.5 shows the shear strength rating for
underground data.
• Moisture Sensitivity Deduction – In the CMRR, the maximum deduction for mois-
ture sensitivity is 15 points. The data sheet used in the Immersion Test is shown in
Figure 2.15 (after Clark & Molinda, 2007). If immersion test has not been done, mois-
ture sensitivity can be estimated visually by examining underground rock exposures.
Underground soft rock mining 71

Table 2.5 Bedding/discontinuity shear strength rating table for underground data

Cohesion

Roughness Strong Moderate Weak Slickensided

Jagged 35 29 24 10
Wavy 35 27 20 10
Planar 35 25 16 10

Depending on the moisture sensitivity of a rock, time is required for the air and
humidity to negatively affect rock strength. Where excavations are only needed
for a relatively short time, applying a moisture sensitivity deduction may be
unnecessary. The CMRR program reports both the Unit Rating and the CMRR
with and without the moisture sensitivity deduction.
Research was conducted by Mark and Molinda (2005) to explore the relation-
ship between the Slake Durability Test (SDT) and the Immersion Test (as detailed in
Figure 2.15). The correlation was less reliable for distinguishing “moderately sen-
sitive” rocks from “severely sensitive” rocks. Table 2.6 indicates how the results
from either test can be used for input to the CMRR.

2.2.2.4 Calculation of CMRR


The calculation of the CMRR is done as follows.
• Obtaining the Unit Rating is as follows:
When using underground data, the equation for calculating the Unit Rating is:

Unit Rating = (UCS Rating + Discontinuity Intensity Rating + Discontinuity


Shear Strength Rating + Multiple Discontinuity Adjustment + Moisture
Sensitivity Deduction)

For drill core data, the equation is even simpler:

Unit Rating = (UCS Rating + Discontinuity Rating


+ Moisture Sensitivity Deduction)

• Obtaining the Thickness-Weighted Average Roof Rating (RRW) is as follows:


The next step in the calculation of the CMRR is to determine the thickness-weighted
average of the Unit Ratings of all the units within the bolted interval.
Table 2.6 Moisture sensitivity classes and rating from both immersion and slake durability tests

Moisture sensitivity class Rating adjustment Immersion index Slake durability index

Not sensitive 0 0–1 100–98


Slightly Sensitive −3 2–4 98–92
Moderately Sensitive −10 5–9 92–80
Severely Sensitive −25 >9 <80
72 Underground soft rock mining

Figure 2.15 Data sheet for immersion test

Consider the following simple example (Mark & Molinda, 2005) that assumes
the roof consists of three units (from top down):
• 2m Sandstone, Unit Rating = 60
• 0.8m Siltstone, Unit Rating = 50
• 0.4m Shale, Unit Rating = 40 (the immediate roof layer)
If length of the roof bolts is 1.8 m, then the thickness-weighted average (RRW) is:

RRW =  ( 0.4 * 40 ) + ( 0.8 * 50 ) + ( 0.6 * 60 ) / 1.8 = 51.1


Underground soft rock mining 73

Note that although the uppermost roof layer was 2m thick, only the lowest 0.6m
(the distance to the top of the bolts) was used in the calculation.

The CMRR is now determined by applying several adjustment factors to the RRW.

• Strong Bed Adjustment (SBADJ) – One of the most important assumptions in the
CMRR is that the strongest bed within the bolted length often determines the
performance of the mine roof. The strong bed’s effect on the CMRR depends first
upon how much stronger it is than the other units. Second, the strong bed must be
at least 0.3m thick before it can provide any additional support, and the amount
of the adjustment is maximum when the bed is at least 1.2m thick. Third, the roof
bolts must obtain at least 0.3m of anchorage in the strong bed for the adjustment
to be considered. Finally, the higher into the roof that the strong bed is located,
the less its positive effect will be. In the original CMRR (Mark & Molinda, 2005),
the SBADJ was determined initially using a table. For improved accuracy and to
facilitate the implementation of the Table in the computer program that NIOSH
produced, based on the author’s research, the following equation was derived
using multiple regression:

SBADJ =  ( 0.72SBD * THSB ) − 2.5  *  1 − ( 0.33THWR − 0.5 )

Where SBD is the Strong Bed Difference, the difference between the strong bed’s
Unit Rating and the thickness-weighted average of all the Unit Ratings within the
bolted interval; THSB is the thickness of the strong bed (m), and THWR is the
thickness of the weak rock suspended from the strong bed (m).
Note that if the strong bed is at the top of the bolted interval, its full thickness
is used in the calculation of the SBADJ (up to a maximum of 1.2m).
• Other Adjustments
• Number of Units – Many workers have indicated that mine roof that con-
tains numerous lithologic contacts is less competent than roof that consists
of a single rock type. When the depositional processes change in geolog-
ical time and distinctly different material is deposited, there is generally,
but not always, a sharp contact between these units. Since gradational
contacts do not weaken the roof, the characteristics of major bedding con-
tact surfaces (cohesion and roughness) should be noted. The maximum
deduction from the CMRR is 5 points when more than four weak contacts
are present.
• Ground Water Adjustment – Ground water is more prevalent in shallow
mines, particularly beneath ground and surface water, but it can also
ingress from abandoned mines or fracturing caused by mining that inter-
sects water-bearing strata. The CMRR maintains the RMR system’s
rating scale, with a maximum deduction for f lowing ground water of
10 points.
• Surcharge – The strength of rocks overlying the bolted interval is only consid-
ered when they are significantly weaker than the rocks within the rockbolted
layer. According to Mark and Molinda (2007), an example was a mine in the
74 Underground soft rock mining

west of the US, where 1.2m of relatively strong top coal was overlain by 6m
of weak, rooted claystone. Due to this, the bolted roof beam needed to carry
some of the surcharge (extra weight) of the incompetent claystone, and the
stability was reduced. The CMRR accounted for the surcharge with a 3-point
deduction.

2.2.3 Coal mine planning


Modern mine planning now relies heavily on computer software to better opti-
mize the mine planning process. This also permits different alternatives to be
rapidly considered and mine short- and long-term plans to be regularly updated
as needed.
The following is a sample of commonly used planning software on soft rock (espe-
cially coal) mines, in random order:

• Surpac
• Vulcan
• Gemcom
• Geovia MINEX
• Minemax
• Minesight
• Deswik
• Carlson

According to Ghosh (SRK, Brisbane, Australia), the following three models are
required for any successful safe and productive mine planning and exploitation:

• The Geological and Resource Model: This geologically defines the orebody and
rockmass and is the starting point for mining engineering design and planning
work. This is the model based on the geologist’s interpretation of drill hole infor-
mation and other relevant surveys. This model allows the engineer to readily vis-
ualize a mine exploitation method.
• The Planning (Engineering) Model: This is important in the planning and
scheduling process. It uses engineering parameters to determine the extent
of the deposit and assess the mineable quantities and qualities of ore. It pro-
vides the practical, mineable interpretation and mine design of the geological
model.
• The Scheduling Model: This is the “time” model used to determine the produc-
tion rate, tonnage, grade and sustainability of the Engineering Model. The “time”
component allows the engineer to calculate what practical production rate and
grade should be generated and maintained over the course of the mine’s life. It
helps identify production and grade problem periods (bottlenecks) so these can
be hopefully eliminated by modifying the Schedule. This then alerts the planning
engineer to make adjustments to designs and plans to accommodate potential
deficiencies in mine production before they become real operational or supply
problems.
Underground soft rock mining 75

2.2.3.1 The geological and resource model


The Geological Model is used to define a rockmass geologically. This includes
determining:

• The stratigraphy
• Defining each rock type and the relevant rock properties
• The regional and local geology
• The hydrology
• The coal seam and its extent and properties
• An idea of the stress tensor

The Resource Model determines the ore tonnage and grade of a geological deposit
using the Geological Model.
Tables 2.7–2.10 give an excellent summary of the stages involved in developing the
Resource and Geological Model (Kennedy, 2013).

2.2.3.2 The planning (engineering) model


This model takes the Resource and Geological Model and produces a mine model that
considers typically of:

• Identifying any areas of the mine lease that are obviously uneconomic to mine
or should not be mined due to sensitive infrastructure, environmental issues that
cannot be readily addressed or valid community concerns.
• A preliminary mining method (or methods) determination, particularly whether
surface or underground extraction is more feasible overall. Typical methods would
be one or a combination of:
• Strip mining
• Highwall mining
• Longwall mining
• Room and pillar mining

The mining method will consider all the rockmass properties, orebody dip, geological
feature locations and orientation and any knowledge of the stress tensor.

• A first production rate estimate that can then be used to establish the life of mine,
equipment suite and manpower levels. This will also depend on the established
market for the coal produced.
• A mine access and mine and support design (or designs) can be prepared.
• A ventilation system will be needed if it is an underground mine, and this can be
designed using any suitable computer software package such as VentSim.
• The surface infrastructure can be designed including any process plant needed
(such as a washing plant) and the location coal stockpiles, waste rock and ash
dams or dumps.
• A first estimate of the revenue, capital and operating costs can be made.
• A first estimate of the mine’s financial viability can be established.
76 Underground soft rock mining

Table 2.7 T
 he logical development of the Resource and Geological Model-the pure geology
component (Kennedy, 2013)

Scoping/Conceptual Design and Operation and


Study Pre-feasibility Feasibility Construction Ongoing Maintenance

Review of Preliminary Review of existing Detailed Review of any


Regional Data review of existing exploration, site-specific regional changes or
exploration geological maps analysis new exploration
mapping, and core logs by mapping, geological
geological maps on-site geologists; maps or core logs
and core logs preliminary site- that may become
specific analysis available
Data Collection Library review, Major structures Ongoing Refine geological
and review of defined; geophysical database
geophysical data geophysical sampling;
if available sampling enhancement
(surface or conducted; of geological
downhole) develop database
database(s)
Geological Preliminary Detailed Geological Review and
Mapping and review of geological teams (mining revise geological
Model existing maps mapping and and external maps, sections
Development and preliminary associated company) and level plans;
geologic mapping; cross sections review and refine 3D model
initial country produced; initial 3D further enhance
rock model model geological
mapping and
associated
cross-sections;
enhance 3D
model and
develop long
sections and
level plans
Geologic Preliminary Basic geological Detailed geological Geological
Assessment geological assessment assessment of assessment
assessment, and review structures/rock revisions
including conducted contacts, made as needed
arrangements alteration (as information
made for mineralization changes)
collecting and deposit
geotechnical trends and review
data for analysis conducted
Mineralogical Limited sampling; Preliminary Detailed Location-specific
Sampling and preliminary mineralogical mineralogical mineralogical
Analyses assessment sampling and sampling and sampling and
analysis; mapping; detailed mapping; revisions
preliminary mineralogical of mineralogical
mineralogical study study study as needed

Using a suitable computer design and planning package, the viability of the mine
design can be optimized. It is also necessary to revisit the Engineering Model regu-
larly as more information becomes available.
Both long- and short-term (more detailed) planning is essential, and these opera-
tions should ideally not be siloed so that they integrate totally and effectively.
Underground soft rock mining 77

Table 2.8 T
 he logical development of the Resource and Geological Model-the drilling and sampling
component (Kennedy, 2013)

Scoping/Conceptual Operation and


Study Pre-feasibility Feasibility Design and Construction Ongoing Maintenance

Drill Hole Spacing Wide spaced Initial infill drilling Infill (closely- Targeted drilling
drilling as of wide-spaced spaced) drilling (as needed) to
estimated from drilling; on a detailed grid clarify any
exploration/ development of pattern as uncertainties;
geological preliminary grid estimated by infill drilling
maps patters; targeted calculated (as needed)
oriented drilling reserve for sampling
for development categories; and testing for
of structural infill-oriented intact rock
model drilling for strength
refinement of
structural model
Underground Review of Drilling if Detailed drilling if Detailed and/or
Drilling existing data accessible accessible targeted
underground
drilling
Sampling Preliminary Geotechnical All feasibility Location-specific
sampling; sampling; test sampling programs samples taken
obtain outcrop pits complete; detailed where needed for
samples where sampling from further refinement
available targeted and/or of design/
infill drilling construction
parameters
Drilling/Assay Preliminary Check of drill Check of drill Ongoing check
Data check of holes holes of drill holes
existing drill (coordinates, (coordinates, (coordinates,
hole data elevations, angles, elevations, angles, elevations,
etc.); check assays, etc.); check assays, angles, etc.);
angled hole vs. angled hole vs. check assays,
vertical hole vertical hole angled hole vs.
comparison; assay comparison; twin vertical hole
flow diagram and hole drilling; assay comparison; twin
development of flow diagram and hole drilling; assay
database validation of flow diagram and
database refinement of
database

2.2.3.3 The scheduling model


The Scheduling Model is the final stage after the Resource and Geological Model
and the Planning Model are basically completed. Scheduling involves determining the
start and end of each of the mining-related activities. This is a complex and involved
process on a mine because of the number of operations that need to be completed
often simultaneously and frequently under space and access constraints.
The goal of scheduling is to arrange all operations to start and stop in a manner
such that safety is ensured, costs minimized and revenue maximized over the life
of the mine. The scheduling process is used to eliminate or reduce any production
bottlenecks so as to maximize productivity and profits (safely).
78 Underground soft rock mining

Table 2.9 T
 he logical development of the Resource and Geological Model-the geotechnical
component (Kennedy, 2013)

Scoping/Conceptual Design and Operation and


Study Pre-feasibility Feasibility Construction Ongoing Maintenance

Pertinent Review of regional Incorporate Review analogical Keep regional


Regional geotechnical data; regional mine design geotechnical
Information analogical mine geotechnical parameters and data and
design cases parameters in regional analogical mine
identified for pre-feasibility geotechnical design cases
comparison and plans and data to compare on file for
benchmarking designs; identify to updated comparison
any differences design for any and reference
between differences or
analogical cases similarities
Assessment and Geotechnical Assessment and Ongoing Refinement of
Database assessment of compilation of assessment and geotechnical
Maintenance advanced initial mine-scale compilation of all database
exploration geotechnical data; new mine-scale
data establish a geotechnical data;
preliminary enhancement to
geotechnical the geotechnical
database database
Intact Rock Use literature Index and Laboratory Sampling on
Strength values laboratory testing on infill/targeted
Determination supplemented testing on detailed drilling and
by index tests sample selected samples; laboratory
on core from from targeted detailed testing
geological mine-scale assessment and
drilling drilling; initial establishment
assessment of of parameters for
lithological 3D geotechnical
domains model
Stress Regional Supplement by No additional
No additional
Measurement compilations; research data data collection
data collection;
world stress from nearby sensitivity study
map mines may be conducted
if numerical
modelling used
3D Geotechnical Consider validity Prepare a draft Enhancement Refinement of
Model of regional of the 3D model of the 3D model 3D model
geotechnical
data as a source
for parameters
in model

According to Roseke (2017) and modified by the authors, the typical scheduling activ-
ities include:

• Plan schedule management


• Activities definition
• Sequencing of all critical activities
• Estimate the required activity resources
• Estimate activity durations
Underground soft rock mining 79

Table 2.10 T
 he logical development of the Resource and Geological Model-the structural component
(Kennedy, 2013)

Scoping/Conceptual Design and Operation and


Study Pre-feasibility Feasibility Construction Ongoing Maintenance

Mapping Collect and Mine-scale outcrop Trench mapping Detailed mapping,


review aerial mapping (if applicable) if needed, in
photos from specific areas
historical times of uncertainty
to present;
regional outcrop
mapping
Stereographic Initial Database Enhancement Refined
and Structural stereographic established of database; interpretation
Domains drafts from for structural advanced of fabric data
outcrop information; stereographic and structural
mapping initial assessment of domains
parameters stereographic fabric data;
assessment of confirmation
data; initial of structural
structural domains
domains
established
3D Structural Initial ground Initial 3D Refinement of Refined
Model proofing structural model 3D structural interpretation
model of 3D structural
model

• Develop the schedule


• Optimize the schedule
• Continually compare plan with actual results and update the scheduling and
requirements for continuous improvements

Some of the activities involved in scheduling include:

• Taking the planning model and basically introducing time to it (e.g., daily, weekly,
monthly and yearly production)
• Defining all the activities required, the critical interactions, how they need to inte-
grate and finding any production bottlenecks, and then trying to eliminate them
• Establishing equipment, manpower and other requirements
• Generating schedules, manpower rosters, Gant charts and other relevant reports

It makes sense to integrate the planning and scheduling by using the same software
package as this maximizes efficiency and reliability.
Effective planning and scheduling help ensure operational and safety excellence.

2.2.4 Mine planning challenges


Most of the challenges associated with the mine planning process are associated with
the accuracy of the data used due to the uncertainty of the geological model. Regularly
revisiting and updating the planning and scheduling models as more information is
available is vital during mine development, exploitation and mine closure.
80 Underground soft rock mining

2.3 Coal mine access and development

2.3.1 Introduction to coal mine access


The method or methods of accessing a new coal deposit depends on the depth below
surface; if shallow, then slopes or declines are preferred, but if deeper, then shaft
access may be needed. This is because conveyor belts are generally more cost effective
and can move more coal than via shaft skips. Adits can only be used if the coal seam
outcrops in a mountainous area.
The cross-sectional area of the access is designed based largely on:

• The planned production


• The ventilation needed (more if it is a gassy deposit)
• The equipment to be used

In addition, the access needs to provide easy travel for all personnel and consumables,
delivery of all services (unless shallow where some can be fitted into a borehole) and
providing water into and out of the mine.
Since mines over time often try to increase production above the original planned
level, it can be advantageous to build in extra ventilation and production capacity.
Many countries regulate that two independent accessways back to surface must be
available to permit workers to escape quickly if necessary, so an alternate mine egress
access is provided in case the main one is lost for some reason.
Depending on the coal seam depth, the main access is provided by a ramp (decline)
from the surface if relatively shallow or a vertical shaft if deep. Ramp access is desir-
able as it provides more ready access, and a conveyor system to the surface is more
productive than coal skips in a vertical shaft. The main access from where men and
materials travel in and out of the mine is also generally used as the fresh air supply for
the mine so must be sized accordingly.
The location of the main access depends on many factors, including:

• Surface topography
• The coal seam location and dip

It is common for ramps to be developed from the surface at about a 15° maximum
angle from the horizontal and flattened for curves if needed. It can vary from 10° to 20°
depending on the efficient incline that the conveyor can handle with the particular coal.
The selection between ramp or shaft on coal mines is different from hard rock mines.
This is because coal seams are tabular and generally relatively flat dipping so ramps can
be equipped with conveyors for coal, whereas trucks are generally used to transport the
ore in hard rock mines, which are often massive or steeply dipping. Conveyors are very
effective in moving coal which is soft and continuously produced unlike hard rock ore
which is produced in a “batch mode” because of drilling and blasting.

2.3.2 Ramp (decline) access


Ramps from surface are still effective for greater depths than for hard rock mines
mainly because conveyor transportation is used, which is very efficient.
Underground soft rock mining 81

The rule of thumb appears to be 500m deep when shafts may become more effective
as the primary access.
Declines (or inclined shafts) are associated with grade constraints and these must
be considered carefully. If a decline is used for trackless haulage, then the maximum
grade recommended is about 8o. However, if used for conveyor belt transportation and
the decline is used by rubber tired trackless equipment on a regular basis, then 15o is the
maximum recommended. If used for conveyor belt haulage only, then the maximum
grade could be 15–25o depending on the material to be conveyed. If equipment has to be
driven up and down the decline regularly, to clean spillage, this will limit the gradient.

2.3.3 Vertical shaft access (refer also to Section 3.3)


(Modified from an unpublished report by R. Missavage, former instructor at Southern
Illinois University Carbondale, 2008)

Shafts are generally used for the following functions:

• To access an ore body


• To transport men and materials to and from underground workings
• For hoisting ore and waste from underground
• To serve as intake and return airways for the mine (ventilation)
• To provide a second egress as required by most countries’ mining law
• Storage of nuclear waste

There are many factors that influence the location and therefore the type of shaft to
be sunk. The shape, size and dip of ore body will mainly dictate where the shaft posi-
tion should be. In addition, host ground conditions and water-bearing structures also
influence the final location of shafts. Generally, there are three types of ore deposits:

• Narrow tabular deposits (ranging from steep to flat dipping such as gold, platinum, etc.)
• Wider tabular deposits (such as coal and potash)
• Massive deposits (diamonds, copper, nickel, iron ore)

A shaft that penetrates from the surface to the coal seam passes through a variety of
overburden soil and rock strata. Consequently, the shaft support requirements vary
considerably from shaft collar on the surface to the shaft bottom, some distance below
the coal seam. The following can be used to develop a procedure for the design of shaft
linings on the basis of the properties and conditions of the soil and rock through which
the shaft passes. Designs for cohesive and cohesionless soil, and elastic-plastic and clas-
tic rock need to be considered if applicable. Clastic rocks are composed of fragments, or
clasts, of fragmented rock. Clastic rocks are generally sedimentary rocks.
The design involves the following steps:

• Estimating the primitive and in-situ stress states (that may change during mining)
along the length of the shaft from surface
• Using the stress conditions to establish the reactive pressure in the shaft lining
that is required to ensure that the shaft will not fail during its useful life
• Designing the shaft lining that will provide the stability required during the shaft
life with an adequate factor of safety (FOS)
82 Underground soft rock mining

2.3.3.1 Shafts developed through overburden soil


Most overburden soils consist of either cohesive (clays and silts) or cohesionless (sands,
gravels, etc.) materials, or a combination of both. The properties of the shaft lining
therefore needs to vary considerably depending upon whether cohesive or cohesion-
less behaviour predominates. It should be noted that severely weathered or altered
rock should also be considered as a soil.
Depending on the nature and depth of any soil cover, the use of ground freezing
can be considered to make the excavation much faster and safer. The peripheral fro-
zen cylinder produced performs the dual functions of groundwater control and earth
support, allowing shaft excavation often without the need for internal bracing and
sheeting.
• Shafts in Cohesionless Soils
The nature of cohesionless soils means that ground support is always required unless
moisture in the soil provides an adequate “apparent cohesion”, but this can be tempo-
rary (it may dry out) so caution must always be exercised. It is therefore always pru-
dent to rather support excavations in cohesionless soils, since any apparent cohesion is
not a dependable soil property. Ground support pressures required for circular shafts
in cohesionless soils above the water table were determined by Terzaghi (1943) to be
(can be converted into SI units):

σh = M σ γ 2 r ( in psf )

Where:

σh is the lateral support pressure (psf)


γ is the soil unit weight (pcf)
r is the shaft radius (ft)
M σ is an empirical factor which is dependent upon the angle of internal friction
of the soil “Φ” and the depth below the surface “h”

h  a + 1   a − ( a − 2 ) + n12 

Mσ =
2 r  2 a   a + n1a +1 


Where:
a
n1 =
a−2

σzz  ϕ 
a= = tan 2  450 + 1 
σrr  2
σrr is the stress in the radial direction from the shaft centre
σzz is the vertical stress
h is the depth of the shaft or excavation (sometimes referred to as z) in ft
r is the shaft radius
φ1 is the angle of internal friction
Underground soft rock mining 83

For values of φ between 30o and 40o, the values of φ1 can be approximated by:

ϕ1 = ϕ − 5 ( in degrees )

Thus “a” and “n1” will have constant values for most sands.
The cohesionless sand is assumed to be above the water table. If the walls of the
shaft are below the water table and the shaft is perfectly sealed (seldom the case or
desirable), the walls will be acted upon by the sum of full water pressure and the
pressure exerted by the sand (based on the submerged unit weight of sand). The sand
pressure is usually negligible when compared with the water pressure. The effects of
water pressure in deep shafts can be generally negated by providing drains that extend
into the soil and rock behind the lining.
• Shafts in Cohesive Soils
Based on Terzaghi (1943), the required ground support pressures can be estimated
from:

σh = γ h − 2C ( psf )

Where:

h is the depth of the shaft or excavation (sometimes referred to as z) in ft


γ is the soil unit weight (pcf)
C is cohesion (psf)

The above expression is valid for deep excavations. For shallow circular excavations,
h < 24r, the number/coefficient “2” can be replaced by “K” whose value is

 h  
K = 2  ln  + 1 + 1
  r  

Where:

h is the depth of the shaft or excavation in ft


r is the shaft radius in ft

The expressions for the required support in cohesive soils imply that the shaft will be
essentially self-supporting down to a critical depth, at which point, the soil pressure
becomes larger than the soil shear strength. Above this depth, ground support would
not be required (in cohesive soils) to maintain the excavation, but a tunnel lining is
suggested anyway to prevent sloughing and unraveling of the shaft walls and localized
failures. This critical cut depth is given by:

KC
hcr = ( in ft )
γ

In Europe, Ostrowski (1972) presented the following values for the required support
pressures in incompetent rocks.
84 Underground soft rock mining

Above the water table:

Sandy formations: σh = 0.5 − 0.6 * 0.434 h ( in psi )

Clay and loam: σh = 1.0 − 1.1 * 0.434 h ( in psi )

Below the water table:

Sandy formations: σh = 1.3 − 1.4 * 0.434 h ( in psi )

Clay and loam: σh = 1.5 − 1.7 * 0.434 h ( in psi )

In relatively short trenches of finite dimensions, such as rectangular shafts, the hori-
zontal ground support pressure may be estimated as:

σh = γ hw + γ′ ( h − hw ) A K a ( in psf )

Where:
γ is the total unit weight of soil (pcf)
γ′ is the submerged unit weight of soil (pcf)
hw is the depth of ground water (ft)
h is the depth of cut (ft)
A is the reduction coefficient
Ka is the active earth pressure coefficient

The value of the reduction coefficient “A” is given by:

1 − exp ( − 2 n K a tan ϕ)
A=
2 n K a tan ϕ

Where:
n is the ratio of depth to length of cut “h/l”

2.3.3.2 Shafts in rock


Shaft support requirements in rock depend on a variety of geotechnical and hydro-
logical parameters.
Stability may be:
• Complete
• Dependent upon geologic structure
• Dependent upon rock strength in-situ stress relationships (or a combination of both)
Complete stability is the condition that typically exists where:
• The rock is strong and massive
• Stresses induced by the shaft are well below the rock mass strength
Underground soft rock mining 85

Even where shafts are completely stable for the life of the mine, a cosmetic lining
may be installed to reduce friction losses in ventilation or to hang shaft hardware for
hoisting, support of utilities, etc. This also helps in case of unexpected seismicity,
weathering or adverse hydrology or geology.
Stress-controlled shaft wall instability (or stability) is governed by:

• Shaft size and shape


• Rock types and properties
• In-situ stress conditions (and any changes due to future mining-induced stresses)

The stability and hence the required lining support pressure (if any) is governed by the
magnitude of the stresses induced in the shaft walls relative to the rock mass strength.
The stresses at the shaft walls are dependent on the shaft shape, the magnitude of the
horizontal stresses and their relative stress uniformity.
A circular opening in a uniform (axisymmetric, i.e., the horizontal stresses are equal
and constant) stress field produces the least stress concentration at the shaft wall. For
a non-uniform stress field (where the orthogonal horizontal stress components are
unequal, but remain constant throughput the stress field), the stress concentration is
least for an elliptical shaft opening which is oriented with its major axis parallel to the
direction of the major horizontal stress field component (oval shaft cross-sections are
therefore good compromises).
Kirsch equations (1898) can be useful for shaft support design of circular and ellipti-
cal cross-sectional shafts. The equations are based on Figure 2.16 (that also shows the
parameters used in the formulae):

P  
 r 2  
 4r 2 3r 4  

σrr =  (1 + k )  1 − 2  − (1 − k )  1 − 2 + 4  cos2θ 
2   d   d d  

P  r2   3r 4  
σθθ =  (1 + k )  1 + 2  + (1 − k )  1 + 4  cos2θ 
2   d   d  

P  2r 2 3r 4  
σrθ =  (1 − k )  1 + 2 − 4  sin2θ 
2   d d  

Simplifying the Kirsch equations, for a circular opening in a “uniform” stress field ( σh ),
the stresses away from the shaft wall can be estimated as:

  r 
2 
σrr = σh  1 −  

 
 b  
  r 2
σθθ = σh  1 +   
  b 
 

Where:
σrr is the radial stress, σθθ is the tangential stress, r is the shaft radius and b is
any radius of interest beyond “r”
86 Underground soft rock mining

Figure 2.16 Stresses around a circular excavation

Once the stress distribution (and any future stress changes) in the rock around a shaft
has been determined, it can be compared to the rock mass strength to determine
whether overstressing of the rock or rock failure is likely. Triaxial strength data may
be used for this purpose. Linear or nonlinear Mohr’s envelopes may be used, depend-
ing on the amount of nonlinearity the rock exhibits and the accuracy that is required.
It must be emphasized that the rock mass strength rather than the rock specimen
strength should be used in this analysis.
The depth of the failed zone is estimated by the intersection of the rock mass
strength curve with the tangential stress curve. If the rock would fail without a lining,
the supporting pressure that must be provided by a lining that would prevent rock fail-
ure from disrupting the function of the shaft may be determined by either assuming
that either the failing rock behaves plastically, or that it behaves in a brittle manner
(typical clastic rock behaviour).
• Pressure support lining design in elastic-plastic rock
The theoretical pressure that must be developed by the lining to support the rock (con-
versely the pressure that the lining must withstand without failing) is:

2  σo   r   σ 
tanβ−1

Pi = σ
 h +    −  o 

tan β  tan β − 1   Ro   tan β − 1 

Where:

Pi is the required supporting pressure


tanβ is the coefficient of passive pressure = (1 + sinϕ) / (1 − sinϕ)
Underground soft rock mining 87

ϕ is the angle of internal friction


σh is the uniform horizontal ground stress
σo is the rock mass unconfined compressive strength (UCS)
r is the shaft radius
Ro is the radius of relaxed or failed zone

It should be noted that although “Ro” is proportional to “r” in the equation, the lining
pressure required is not a function of the shaft radius “r”. The primary variables to be
considered are tanβ, σo and σh and for σo > 2σh , failure will not occur. The controlla-
bility factor is the difference between σo and σh .
Note that as the ratio “ σo / σh ” approaches 2.0, the lining pressure approaches zero.
• Pressure support lining design in brittle (clastic) rock
The theoretical pressure that must be developed by the lining in clastic rock is:

 C  r tanβ−1 C m
Pi =  m + σh (1 − sinϕ) −
 tanϕ  Ro tanϕ

Where:

Pi is the radial pressure between outside of lining and rock


ϕ is the angle of internal friction
(
C m is the rock mass cohesion = σo / 2 tanβ )
Some points worth noting about shaft lining for rock are:

• As the width of relaxed zone approaches zero, the “r / Ro ” term approaches 1. For
this case, the lining pressure is a function of “ σh ” and “ ϕ ” only.
• As σo approaches 2σh , the width of the relaxed zone approaches zero, which and
consequently the “r / Ro ” term again approaches 1.
• The computed lining pressure is independent of the shaft radius “r” and is directly
proportional to the uniform horizontal stress “σh ”. The lining pressures are func-
tions of variables “σo / σh ” and “ϕ ” as mentioned earlier.

It is suggested that both the elastic-plastic and clastic solutions be calculated to deter-
mine which of the two cases is less stable. The higher of the two pressures calculated
should be used for the lining design, unless it is determined that the less conservative
value can be actually justified.
• Design of shaft lining
The lining design phase is a structural engineering task. If a concrete lining is planned
as is common in the United States, Lamé’s (1837) equation for thick-walled cylinders
can be employed to determine the required lining thickness:

 fC / FS 
t =   a
 fC / FS − 2 Pi 
88 Underground soft rock mining

Where:

fC is the compressive strength of concrete


FS is the safety factor selected for the lining
Pi is the uniform pressure between lining and rock; a is the inside radius of lining
t is the required lining thickness

Some pertinent facts about lining design:

• fC is the UCS of the concrete. Since the lining will be exposed to a three-
dimensional state of stress, the required thickness may be overestimated due to
the strength increase of materials under triaxial conditions. This conservatism is,
however, offset by the fact that the horizontal stress field is never precisely uni-
form, and some localized bending stresses may be present in the lining. It is worth
retaining this conservatism to cover any unknown factors that may increase the
lining loads, or create localized adverse stresses.
• If the excavation extends beyond the payline (the theoretical boundaries of the
shaft), the overbreak zone will normally be filled with concrete during lining
placement, increasing the effective thickness of the lining. Based on previous
shaft lining experience (researcher unknown), the overbreak can be empirically
assumed with some confidence to be:

Overbreak = 6.6728 r + 0.0307 ( in inches and then converted to SI units )

Where:

r is the shaft radius in ft

Due to this additional thickness needed to be filled with concrete (into any overbreak)
to create a constant and smooth final finished shaft to minimize ventilation losses, the
safety factor on the lining thickness can be possibly minimized. A safety factor of 1.50
is considered appropriate and is in compliance with the American Concrete Institute
building code.

• Shaft support with rock bolts, steel sets and shotcrete

Peck (1969) developed a set of guidelines for shaft support using rock bolts, steel sets
and shotcrete. These guidelines were developed by a combination of an “analytical
approach” and an “empirical approach”. In general, the requirement for support in
shafts is less than in tunnels where the stress conditions are less uniform and the rock
is more likely to be further disturbed by mining activities.
The self-supporting capabilities of rock are therefore enhanced by the support. The
final design of a rock bolt, steel set or shotcrete support for a given shaft should be based
on consideration of the possible mechanisms of failure associated with specific conditions
around the shaft and the manner in which the supports might enhance the stability.
The attached guidelines apply to shafts with diameters between 4.6m and 7.6m.
For smaller diameters, somewhat less support is required, whereas considerably more
support would be required for larger diameters.
Underground soft rock mining 89

The length of the rock bolt required is determined based on consideration of the
probable size of the loosening rock blocks. A rock bolt length of 1/2 to 1/3 the shaft
diameter, with a maximum length of 2.4–3.0 m, is recommended (as a first estimate).

2.3.4 Shallow shaft protection pillar design


At shallow depths, several empirical methods can be used to determine a first estimate
of the required shaft pillar dimensions. They all use a linear relationship such that the
shaft pillar size increases proportionally with depth increase.

D = H/5

D = 150 + ( H/5 )

D = 0.72 ( H )

Where:

D is the shaft pillar diameter in m.


H is the depth below surface in m.

The final design should be determined preferably using an appropriate and suitably
calibrated computer simulation package.

2.3.5 Main entry development


Once the coal seam has been accessed from surface via a ramp or shaft, main entries
are developed on the coal seam in order to access the entire coal mine lease. The lay-
out can then be used for either longwall mining or room and pillar mining (usually in
shallower coal seams only).
It is usual to develop at least three entries to give better security in the long term
should one fail for some reason.
The main entries (also referred to as main roadways) have chain pillars left between
them with factors of safety around 2.0. They are protected from the coal mining (using
room and pillar or longwalling) by barrier pillars, examples of which are given in
Figure 2.17 (Peng, 2008) and Figure 2.18 (Pongpanya et al, 2017).
In coal mines, the main entries are developed in the coal seam using a continuous
miner. The development and the later production are achieved using suites of equip-
ment to maximize the safe production and contain costs. As a minimum, a typical
development suite of equipment would be a continuous miner, a rockbolter and two
ramcars or shuttle cars (to remove the coal to the conveyor and bring in equipment
and consumables).
A ramcar (or hauler) consists of two articulated parts. One is for the driver com-
partment and power system (battery or diesel) and the other a large flat bucket with a
back plate (pusher) that can eject the coal using a hydraulic cylinder.
A shuttle car is electrically powered commonly using a trailing cable and reel from
an electrical transformer. The vehicle is rubber tyred, all-wheel drive and has a scraper
90 Underground soft rock mining

Figure 2.17 Main entry protection with barrier pillars in room and pillar mining (Peng, 2008)

conveyor in a central trough for loading coal from the continuous miner and then
discharging it onto the conveyor system via a breaker (to crush any oversize product)
to surface. Shuttle cars can also be battery powered or even diesel powered although
diesel is undesirable due mainly to nano-diesel particulates that are a potentially seri-
ous health hazard.

Figure 2.18 Main entry protection with barrier pillars in longwall coal mining (Pongpanya et al., 2017)
Underground soft rock mining 91

Shuttle cars are easier to operate because they can load and discharge either side,
whereas a ram car can only load and discharge from one side.
Flexible conveyor trains or bridge conveyors can also be used instead of mobile cars
to move the coal to the conveyor system to surface. Whilst both these systems offer con-
tinuous production matching output from the continuous miner but are more involved
systems with many working parts and can be relatively costly to operate and maintain.

2.3.6 Main access protection


All access from surface to the mining panels must be adequately supported for the
long term, often for the life of the mine. It is therefore critical that they are adequately
supported and protected.
Main entries are protected by barrier pillars that provide the necessary long-term
protection to the entries as mining longwall or room and pillar panels are removed.
Pillars tend to become infinitely strong at a width-to-height ratio typically more
than 6. If the immediate roof or floor is potentially weak, then much larger pillars may
be required.

2.4 Room and pillar mine design

2.4.1 Pillar design considerations

2.4.1.1 Main uses for pillars


In any underground mining operation, two different categories of pillars are encoun-
tered: support pillars and protective pillars. The differences between the two catego-
ries of pillars are often not clearly visible, and there are indeed a number of instances
when pillars fulfill both requirements. However, there can be a number of significant
differences in the two types of pillars:

• Support Pillars – Support pillars can be further divided into two classes: pillars
that provide local support and pillars that provide regional support. Pillars often
provide both local and regional support. A good example of this is a conven-
tional room and pillar mining layout that has been designed at a high safety fac-
tor. Local-support pillars have often only temporary use and are extracted (or
partially extracted) once they have fulfilled their purpose. One of the interesting
aspects of local-support pillars is that their useful function is often limited to
the time when actual mining takes place in their immediate vicinity. Subsequent
failure of these pillars can take place provided the mode of failure is slow and
gradual. The concept of yielding support pillars falls into this category. Barrier
and wide inter-panel pillars are typical examples of pillars that provide regional
support.
• Protective Pillars – In the course of mining, it often becomes essential to protect
underground and surface structures from the adverse effects of mining. One of the
practical means of achieving this is to leave portions of the coal seam un-mined
to form protective pillars, or consider other methods (or combinations) such as
backfilling. The design criteria for these pillars depends largely on the nature of
the structure that needs to be protected. In case of surface structures, the design
92 Underground soft rock mining

criterion is based on the magnitude of the surface movements and strains that can
be tolerated by the structure. In the case of underground structures such as bunkers,
pump stations, service excavation, etc., it is usually the magnitude of the stresses
(and mining-induced stress changes) that determines the size of protective pillars.

2.4.1.2 Basic pillar design principles


The classical approach to the design of engineering structures is to determine the
strength of the structure and compare it with the stresses acting upon the structure. The
purpose of the design is to change the dimensions of the structure or its material reduced
by a safety factor. The safety factor makes allowance for variations in the strength of the
material and uncertainties concerning the magnitude of the maximum stress.
The dimensions of a pillar obviously have a significant effect on the strength and
the post-failure performance of a pillar as shown in Figure 2.19. In addition, the depth
(the stress tensor) at which pillars are used is also important as the design approach is
different. Rigid or yielding pillars may be required and the post-failure behaviour of
pillars governs whether they fail in an uncontrolled (desirable) or uncontrolled (usu-
ally suddenly and violently) manner. This is illustrated in Figure 2.20.

STRENGTH OF COAL PILLARS

Because of the complex loading conditions and the difficulties in determining the in-situ
strength properties of the pillar material, a somewhat simplified approach is generally
adopted in the design of mine pillars. Instead of determining the maximum stress that
acts on the pillar, the average pillar stress is commonly used. In addition, rather than
using the strength of the pillar material, empirical formulae that predict the strength of
the whole pillar have been developed. These formulae suffer from the disadvantage that
they are valid only for the conditions under which they have been derived.

FACTORS AFFECTING PILLAR STRENGTH

Since the strength of a pillar is a function of the material strength and the distribution
of stresses in the pillar, it is important to examine the factors that influence the latter.
Probably the most important factors in this respect are the vertical and horizontal

Figure 2.19 The stress-versus-strain behaviour of pillars at different width-to-height ratios (Das, 1986)
Underground soft rock mining 93

Figure 2.20 T
 he post failure performance of pillars (a stiff load line implies a more violent/brittle
failure than a soft loading line) (Ozbay & Roberts, 1988)

components of the stress vector at the contact between the pillar and the surround-
ing rock (the floor and the roof contacts). The distribution of these components and
their magnitude are influenced strongly by the friction and cohesion along the contact
plane and by the deformation characteristics of the pillar material and the surround-
ing rock strata. Hence any method of predicting pillar strength must take into account
the properties of the pillar material and the surrounding rock strata, as well as the
nature of the contact surfaces.
An important feature of pillars is that their strength tends to increase with their ratio
of width, w, to height, h. This increase in strength is due to the lateral confining effect
imparted to the pillar by the lateral component of the contact stress. The effect of the
width-to-height ratio on the strength of pillars can be expressed by following relation:

Pillar strength = C * h α * wβ

Where:

C is the compressive strength of the pillar material (refer to Table 2.11 for examples
of the constants in different coal seams).
α and β are appropriately chosen constants
An examination of the formulae for pillar strength published by various authors
(Salamon & Munro, 1967 and Iannacchione et al., 1992) shows that the value of con-
stants α varies from −1.0 – for laboratory pillar tests to −0.83 for moderately large
model pillars. The value of constant β was found to be about 0.5 in all the model pillar
tests, while a value of 0.46 was obtained from the statistical analysis of a large number
of case histories. Note that the effect of the width of a pillar on its strength tends to be
independent of the size of the pillars, while the effect of pillar height is greatest in the
case of small pillars but tends to decrease with pillar size. Table 2.11 gives examples of
the constants in different coal mining areas (Ianacchione et al, 1992).
Pillar shape is important to the pillar performance, and its influence is discussed
later.
94 Underground soft rock mining

Table 2.11 Exponent estimation for the pillar strength constants (Iannacchione et al, 1992)

Coal sample site α β

South African coal in-situ failures –0.66 ± 0.16 0.46


Pittsburgh coal tests –0.83 0.50
West Virginia coal tests –1.0 0.50

STRENGTH OF ROOF AND FLOOR

One important aspect in room (bord) and pillar mining is the strength of the roof and
the floor. Weak roof and floor strata in the immediate vicinity of the coal seam can
affect the following in room and pillar mining:

• The local roof support in the rooms


• The working conditions and the use of mechanized equipment
• The long-term stability of the overall room and pillar system (i.e., pillar punching
is possible)

As far as local roof control in room and pillar workings is concerned, the problems
tend to increase if:

• The density of the bedding planes in the immediate roof strata increases
• The strength of the rock composing the individual beds decreases
• The horizontal stresses are high

In principle, there are two ways of improving the behaviour of the immediate roof:

• By increasing the apparent strength of the roof


• By reducing the induced stresses in the roof

The former can be achieved by the introduction of support, whereas the latter con-
cerns mine design, excavation orientation and layout. The most important parameters
that control the magnitude of the induced stresses are the size of the pillars and the
room width. In general, the magnitude of the induced stresses decreases with increas-
ing pillar size and decreasing room width. In laminated or bedded roof strata, marked
improvements in the quality of the roof can often be achieved with relatively small
reductions in room width.
The presence of a relatively thin, soft shale band or coal seam in the roof or floor
strata of the seams that are being worked can have a significant influence on the
quality of the immediate roof or floor, even if these are composed of relatively com-
petent sandstone. An example of excessive floor heave caused by a very soft shale
band (sh) beneath an otherwise competent sandstone bed (st) is shown in Figure 2.21
(Wagner, 1980). Phenomena of this kind are usually confined to relatively deep-lying
coal seams. This type of floor heave is often observed in secondary pillar-extraction
sections where the situation is aggravated by the high abutment stresses that act on
the pillars in the stooping line (the line between partially extracted pillars and intact
pillars). In the case of very narrow seams, floor heave can be so severe that it becomes
a serious production problem.
Underground soft rock mining 95

Figure 2.21 Example of excessive floor heave caused by weak immediate floor strata (Wagner, 1980)

2.4.1.3 Pillar load


The prediction of pillar load is one of the principal problems in the design of
pillar layouts. Until recently, the only theoretical method of determining pillar
load was based on the tributary-area theory. This theory applies to horizontal
workings of which the lateral extent is very large in relation to the depth below
surface. Under these conditions, the average vertical pillar stress is by the fol-
lowing equation:

 A
pm = γ H  
 Ap 
 

Where:

γ is the mass of rock per unit volume


H is the depth of seam below surface
Ap is the area of the pillar
A is the corresponding tributary area as shown in Figure 2.22.

Figure 2.22 Tributary area and pillar area


96 Underground soft rock mining

The average pillar stress can also be expressed in terms of the extraction ratio:

A − AP
e=
A

That is:
γH
pm =
1− e

It is important to note that the pillar stress, pm, calculated from equations above, rep-
resents the upper limit of the average pillar load in any horizontal working of constant
height supported by uniform pillars.

2.4.2 Room and pillar panel design

2.4.2.1 Tributary Area Theory


The theoretical pillar stress is usually obtained by either computer analysis or by using
the Tributary Area Theory (as shown in Figure 2.23). This theory ignores the effects of
abutments and hence is only really applicable where the span of the mining is at least
equal to the depth of the mining below surface. Should this not be the situation, the
pillar loads obtained will be substantially higher than the actual pillar loads, resulting
in a very conservative approach.
Tributary Area Theory is used as follows (assuming a regular pillar layout).
The total tributary area supported by a pillar is:

Ap = ( w1 + b1 )( w2 + b2 )

The area of the pillar is:

A = ( w1 * w2 )

Figure 2.23 The basis of Tributary Area Theory


Underground soft rock mining 97

Assuming the depth of the pillars to be h (in kms) and the average rock density to be ρ,
the overburden stress is:

σv = ρ g h ( in MPa )

The pillar load (stress) is therefore:


 ( w + b )( w + b ) 
σ p = ( ρ g h )  1 1  ( in MPa )
2 2
 
 ( w1 * w2 ) 

The percentage extraction (e) is equal to:


   
  ( w1 * w2 )  
e =  1−   * 100 
 ( w1 + b1 ) * ( w2 + b2 )
 
  

2.4.2.2 Pillar strength


The strength of pillars (Ps) has been found by back analysis to be:

Ps = k ( w 0.46 /h 0.66 ) (1) (after Salamon & Munro, 1967)

Ps = k ( w 0.50 /h 0.77 ) (2) (after Hedley & Grant, 1972)

Where:
k is a constant depending on the rock type
w is the pillar width (assumed square)
h is the effective pillar height
Equation (1) was based on research into coal mines and k was found to be 7.2 MPa for
South African coal.
Equation (2) was based on research into granite rocks in Canada and k was
found to be 133 MPa (Hedley and Grant, 1972). Back analyses on typical quartz-
ites in gold mines in South Africa by Spearing (internal mine report, 1988) found
that k was between 110 and 140 MPa.
A detailed study of the failure process of coal pillars showed that the failure
commences at the circumference of the pillar and migrates inwards. On the basis
of these observations, it was suggested that the ratio of the area, Ap, to the circum-
ference, C, of a coal pillar has a strong influence on the pillar strength (Wagner,
1980).
Irregular pillars (particularly rectangular pillars) therefore can be approximated as
square pillars by calculating an equivalent square pillar with an effective pillar width
(after Wagner, 1980):

Effective pillar width = weff = 4A/R


98 Underground soft rock mining

Where:

A is the area of the irregular pillar


R is the circumference of the pillar

2.4.2.3 Pillar factors of safety


The greatest hazard associated with pillar mining is the potential of violent pillar
failure and collapse. Sudden failure is not always preceded by pillar spalling, and once
the failure starts, it is virtually impossible to stop it.
An acceptable FOS between 1.5 and 1.6 has been found safe in room and pillars
workings on coal mines. Reliability concepts are, however, useful for most pillar
design applications.

• The use of reliability concepts in pillar design

The strength of pillars depends on:

• Rock strength
• Pillar dimensions

For a standard pillar design approach, these parameters are taken as constant,
whereas in practice, the parameters vary considerably (e.g., the pillars are often cut
into variable sizes and the height of each pillar is seldom the same due to stoping width
variations).
Similarly, the stresses on pillars are assumed to be the same, but this is obviously also
not the case in practice due to the different mining spans and the varying overburden.
Reliability concepts consider these variations that can be of major concern in the
FOS of pillar design, and a simple illustration is given in Figure 2.24.
In Cases 1 and 2, the FOS is the same and equal to:

FOS = Average pillar strength/Average pillar stress

Figure 2.24 Pillar reliability examples


Underground soft rock mining 99

Figure 2.25 Pillar failure examples (Monsalve et al., 2020)

However, in Case 1, the variability of the pillar strength and stress is quite large and
hence the possibility of pillar collapse exists (the shaded area shown in the Figure). In
Case 2, the variability of the parameters is relatively small and hence the possibility of
pillar collapse is insignificant.
It is clear from the above very simple example that variability must be considered in
any pillar design. This includes the actual variation in pillar dimensions and shape as
well as rockmass and geology variations.
Pillar failures can be broadly defined by either structural or stress as shown in
Figure 2.25.
In order to undertake such an investigation on a mine, the following data must be
gathered:

• The unconfined compression strength (UCS) of the orebody itself at numer-


ous locations to determine an average value and a standard deviation (Young’s
Modulus and Poisson’s Ratio would also be useful)
• The stratigraphic column at various locations above the orebody to determine the
average overburden density
• Actual pillar dimensions
• Mining dimensions and spans (room width, etc.)

The use of such data can be beneficial when designing pillars by identifying potential
instabilities that the use of average stresses and strengths would not reveal. This tech-
nique can therefore be used to determine the actual factor of safety required.
There are many pillar strength formulae, derived either empirically, analytically or
numerically. These formulae are compared in Figure 2.26, which can be divided into
the three categories as follows (Michaud, 2017):

• Empirical formulas, comprising those of Bieniawski, Sheory and Obert-Duvall,


who take a middle road, implying a linear increase in pillar strength.
100 Underground soft rock mining

Figure 2.26 Comparison of different pillar strength formulae and models (Michaud, 2017)

• Analytical models, such as the Salamon-Munro, Holland-Gaddy empirical formu-


las and Barron’s analytic formula, indicate the core stress, and the pillar strength,
tend towards some maximum and limiting value. as shown in Figure 2.26.
• Numerical models, comprising Hsuing-Peng and Kripakov, Wilson’s formula and
Salamon’s “squat” formula, indicate that the pillar strength increase exponen-
tially as the pillar width is increased. These theories all assume that confinement
grows rapidly within squat pillars, allowing the core stress to build indefinitely.

2.4.3 Barrier pillar design


Barrier pillars are generally used to protect the main entries in longwall and room and
pillar mines and also to isolate room and pillar panels (sections) from each other. This
is to limit the possibility of massive pillar runs from one panel to another.
Such pillars are also essential when mining close to other mines, especially consid-
ering the potential for water and/or gas inrushes.
A rule of thumb is that if a pillar is at least six times the seam mined height, then
assuming component roof and floor, it is infinitely strong. This is useful as a first
rough estimate of the pillar width. Clearly under-designed pillars can lead to massive
potential hazards whilst over-designed pillars lead to a loss in revenue.
In typical room and pillar mines, a factor of safety for in-panel pillars of 1.6 is a
good target; for most barrier pillars, it is about 2.0.
Suitable computer aided design, if correctly modelled and validated, is the most
reliable design method, especially because sensitivity analyses can be quickly under-
taken. A good reliable empirically designed method (that can be used within the data
Underground soft rock mining 101

parameters used to create the method) is the Analysis of Coal Pillar Stability (ACPS).
According to Mark and Agioutantis (2018), this method combines the Analysis of
Longwall Pillar Stability (ALPS), the Analysis of Retreat Mining Stability (ARMPS)
and the Analysis of Multiple Seam Stability (AMSS) design methods.
In the US, longwall barrier pillars ranged between 152m and 183m wide, but with
the application of effective computer software packages, they have been safely reduced
to 61–92m (Peng, 2020).
Continuously monitoring and observing the behaviour of pillars is still essential.

2.5 Longwall mining

2.5.1 Longwall panel design


(Largely based on an unpublished report by Pillalamarry and modified by Spearing,
2009)
The principle behind the longwall mining system is that the roof is supported only
immediately adjacent to the worked seam, keeping the blocks of immediate roof
clamped securely from the face to the rear of the supports. At the rear of the sup-
ports, the roof is allowed to fail and collapse, with the swell of the broken material
eventually establishing equilibrium in the strata behind the mined area, carrying an
increasing reactive load.
There are two distinct types of longwall mining. The longwall retreating system
is the most common in the United States, South Africa and Australia. The longwall
advancing system was common in the remaining industrialized countries such as
Britain, Germany, Poland and the Soviet Union, but this has changed due to the over-
all advantage of the retreat longwall method. The methods are shown in Figure 2.27.
The only advantage of the advancing longwall method is:

• Production can start earlier as the gates (entries) don’t need to be complete around
the proposed panel.

The main advantages of the retreat longwall mining method are:

• The geology of the whole longwall panel is well understood before it starts as all
the access and airways around the new panel are completed first.
• Ventilation and methane control are easier and more effective.
• All entries are completed before longwalling so there is less operational conges-
tion and it is easier to manage.
• The gates (entries) are easier and less costly to support as with retreat mining, the
used portion is surrounded by solid coal (around the longwall face position, the
stresses are very high but this is common in both methods).

The longwall retreating system comprises a mining panel, where the roadways that
form the head entry (the end at which mineral is delivered from the face conveyor to
a stage-loader conveyor) and the tail entry are pre-driven before the longwall units
starts production. In the longwall advancing system, the longwall unit makes the head
and tail entries as the face advances, utilizing various techniques to prevent the entries
from being destroyed as they are left in a consolidating gob area. Figure 2.28 show the
typical longwall mining workings.
102 Underground soft rock mining

Figure 2.27 Longwall retreat and advanced mining systems (Peng, 2020 )

As shown in Figure 2.28, at the rear of the supports, the roof is allowed to fall and
break up, with the swell of the broken material eventually establishing equilibrium in
the strata behind the mined area.
Longwall mining is the safest, most productive and generally has the highest extrac-
tion (typically about 80%) of any other underground coal mining method. It does,
however, induce significant ground failure and subsidence as the overlying strata col-
lapses (gobs or goafs) behind the longwall shields.
Figure 2.29 (Peng, 2020) shows the typical layout for a longwall panel with the main
entries and pillars.
The advantages of longwall mining compared to other methods of pillar extraction
are:

• Permanent supports are only needed in the entries and in the installation and
recovery chambers (rooms). Other roof supports (longwall chocks or shields on
modern longwalls) are moved and relocated with the face equipment.
Underground soft rock mining 103

Figure 2.28 Section through a typical longwall mining panel


Source: Mine Subsidence Engineering Consultants (MSEC), 2007, Conference

• Resource recovery of the longwall panel is very high – in theory, 100% of the block
of coal being extracted, though in practice there is always some coal spillage or
leakage off the face haulage system lost into the goaf, especially if there is signif-
icant water on the face.
• Longwall mining systems are capable of producing significant outputs from a sin-
gle longwall face – 8 million tonnes per annum or even more, depending on the
seam height and longwall panel width.

Figure 2.29 Longwall mining panel layout (Peng, 2020)


104 Underground soft rock mining

• When operating correctly, the coal is mined in a systematic, basically continuous


and repetitive process which is ideal for strata control and for associated mining
operations.
• Labour costs/tonne produced are relatively low.

The disadvantages are:

• There is a high capital cost for equipment, though probably not as high as first
appears when compared to the number of continuous miner units which would be
required to produce the same output.
• Operations are very concentrated (so if there is a serious problem, total produc-
tion can be lost until it is resolved).
• Longwalls are not very flexible and are “unforgiving” – they do not handle seam
discontinuities (vertical displacements) well; gate roads have to be driven to high
standards or problems will arise; good face conditions often depend on produc-
tion being more or less continuous, so problems which cause delays can com-
pound and become major events.
• The unforgiving nature of longwalling means that experienced and well-trained
labour is essential for successful operations.
• Advance developments of the entries needed to start the next longwall must be
carefully scheduled and planned so as not to delay the start of the next longwall
panel.

2.5.2 Stresses around a longwall panel


A longwall gives rise to major redistribution of strata stresses and considerable defor-
mation in the gates, particularly the tailgate. The roof behind a longwall is normally
allowed to cave and cause subsidence of the overlying strata and eventually give rise
to a trough depression at the surface above it. Consequently, reestablishment of strata
loading within the caved region behind the longwall depends upon effective caving
and subsequent re-compaction (from the bulking of the goaf). Under these conditions,
the resulting side abutment pressure and coal pillar strength need to be considered for
a panel layout. The results of this analysis then lead to the estimate of the coal pillar
entry sizes required for support during the mining of the longwall panel, and the sub-
sequent one, where the maingate is re-used and becomes the tailgate.
The side-abutment pressure develops at the chain pillar of a wide longwall panel as
well as at the side of an entry as a result of removing the coal which counterbalanced
the forces acting on the opening boundary (Figure 2.30). It would be expected that the
magnitude of the side-abutment pressure for a wider panel is much higher than that
for an entry of the conventional room and pillar mining. The estimation of the pres-
sure requires the study of rock mass movement around the wide working face.

2.5.3 Longwall equipment


A major decision to be made is the size of longwall blocks (width and length). Modern
longwalls involve a large number of pieces of equipment (with many components
weighing 30 tonnes or more); the process of recovering the equipment from a com-
pleted block, transporting it to a new block and then installing it in the new block
Underground soft rock mining 105

Figure 2.30 Pressure redistribution around longwall mining panel (Alehossein & Poulsen, 2010)

(often with some of it being taken out of the mine for overhaul or rebuild on the way)
is a very major operation. Apart from the direct cost involved, production and hence
income is zero during this longwall move period. Bigger longwall blocks will enable
the number of relocations to be minimized; however, there are limiting factors par-
ticularly to the size width of longwall blocks:

• The wider the face, the more power is required on the face coal haulage system
(mainly the armoured face conveyor – AFC). The greater the power, the larger
the physical size of the drive units (usually there is a drive unit at both ends of the
face). The drive units have to fit into the excavation and allow room for access past
them, for ventilation across the face and for some degree of roof-to-floor closure.
In addition, the greater the power, the larger (and therefore heavier) the chain on
the face conveyor – these chains have to be manhandled on the face at times and
there are practical limitations as to the size of the chain. Another major constraint
to the longwall width is that the AFC must have the strength such that it can be
started fully laden with coal. The AFC chain and drive sprocket must be strong
enough to handle this without being damaged.
• In some longwall installations, the heat created by the high-power haulage drives
may become a factor.
• Both face width and length may be governed by limitations created by lease
boundaries, seam discontinuities or planar variations, already existing mine
development and/or ventilation capacity.
• The ability of the mine to develop new longwall blocks in time so that longwall pro-
duction continuity is not adversely impacted. Condition of equipment – changing
out some items for overhaul or replacement during the life of a longwall block can
be very problematic and inconvenient and is best done during a relocation.
106 Underground soft rock mining

Two parallel sets of tunnels/gates (development panels) are cut into a seam with
continuous mining machines, then connected with a set of crosscuts (especially for
ventilation) at the far end. Mining then proceeds back toward the entrance (in retreat).
Longwall mining employs some very productive machines. Large drum shearers, for
example, run on a track (part of the AFC structure) along the length of the coal face
for typically some 180–300m. The shearers cut the coal, which falls onto an AFC (con-
veyor system or pan) that also runs the entire length of the face. The longwall crew
works under a roof of heavy steel supports called shields. As the coal is removed,
shearers, pans and shields are moved forward systematically and automatically by
hydraulic rams. The roof is allowed to collapse behind the shields. In addition to
increasing production, longwall mining is considered to be safer because the miners
work under strong steel shields.

2.5.3.1 The Armoured Face Conveyor


The longwall shearer moves along the length of the face on a rail system that is part of
the AFC (Figure 2.31) driving through a chainless haulage system (which resembles a
ruggedized rack and pinion system especially developed for mining). The shearer moves
at a speed of 10–30m/min depending mainly on cutting conditions and seam height. The
AFC on which the shearer sits is placed in front of the powered roof supports, and the
shearing action of the rotating drums cutting into the coal seam disintegrates the coal
and deposits it on the AFC with the help of the cowling fitted to the shearer drum.
The AFC is driven using a drivehead as shown in Figure 2.32. It consists of:

• Drive frame
• Drive units

Figure 2.31 The components of an AFC


Reproduced with kind permission from Komatsu Mining Corp.
Underground soft rock mining 107

Figure 2.32 The drivehead assembly and the chain sprocket


Source: Ostroj AS, Czech Republic

• Electric motor
• Gearbox
• Fluid coupling
• Chain sprockets

At the maingate, the coal is usually reduced in size in a crusher, and loaded onto the
first conveyor belt by the beam stage loader (BSL) an example of which is shown on
Figure 2.33. Stageloaders typically include the following components:

• Drives, systems and frames


• Crusher
• Mobile tail pieces

2.5.3.2 The longwall plough (plow) and shearer


Coal is removed from the face using either a plough or a shearer.

• Plough

Figure 2.33 A stageloader


Reproduced with kind permission from Komatsu mining Corp.
108 Underground soft rock mining

Figure 2.34 A plough (plow) on an AFC


Photographs provided courtesy of Caterpillar. © Caterpillar Inc. All rights reserved.

The plough is equipped with multiple scrapers that remove coal mechanically at
depths per pass of typically 5–20 cm that falls onto the conveyor and is shown on
Figure 2.34. The plough is mounted on a chain conveyor and the blade moves with
cables. The plough height is fixed at the height of the longwall coal seam. The plough
is pushed against the coal face using hydraulic cylinders against the AFC. The plough
has no rotating parts and is most suitable for soft coal.
The main advantages of ploughs over shearers are:

• Cheap
• Simple (no moving parts on the cutting machine itself)
• Relatively low dust is created
Underground soft rock mining 109

• Able to keep exposed roof area very small (but a large number of chock move-
ments would be required to maintain this)
• Though only a small web is taken, in the right conditions, production rates can be
comparable to a shearer as the plough is operated at a relatively fast speed along
the face

The main disadvantages are:

• The cutting height is fixed


• The ability to cut harder rock intrusions effectively is limited
• Ploughs have a limited maximum coal seam height due to stability issues
• The pull chain is exposed and creates a potential hazard

The plough is seldom used now and the shearer is much more common.

• Shearer

A typical shearer consists of the following components (Figure 2.35):

• Drum or cutting head


• Ranging arm
• Self-haulage system
• Cowls
• Cable handler
• Controls

The coal is cut from the coalface by the shearer (power loader). This machine can
weigh 75–120 tonnes typically and comprises a main body, housing the electrical func-
tions, the tractive motive units to move the shearer along the coalface and pumping
units (to power both hydraulic and water functions).
A device on the shearer used to improve coal loading is a “cowl/cowling” as shown
in Figure 2.36. This is a curved steel plate mounted on supporting arms (that can

Figure 2.35 A longwall shearer (without cowlings)


Photograph provided courtesy of Caterpillar. © Caterpillar Inc. All rights reserved.
110 Underground soft rock mining

Figure 2.36 A cowling on a shearer, with shields and the AFC


Photograph provided courtesy of Caterpillar. © Caterpillar Inc. All rights reserved.

rotate) so that it sits behind a cutting drum to prevent coal passing from the back of
the drum into the area where the web has been cut; the coal is therefore being forced
to move sideways towards the AFC instead.
The cowl is fitted with a motor and drive mechanism so that the cowl can be
rotated from one side of the drum to the other when the cutting direction is
reversed.
The shearer (Figure 2.37a) obviously needs considerable power, and the power sup-
ply must be continuously available along the width of the longwall. This is provided
using a cable and hose handler often called a “Bretby” (Figure 2.37b). Fixed cables
and hoses are run in a services tray attached to the back plate of the AFC from the
services supply end of the face (normally the maingate) to mid-face. At this point, the
shearer trailing cables and hoses, enclosed within the Bretby, are attached to the fixed
cables and hoses, with the Bretby then being run in a cable trough attached to the
AFC back plate (above the services tray) to both ends of the face as the shearer moves
from end to end.

2.5.3.3 The longwall shields


A number of hydraulic jacks, called powered roof supports, chocks or shields, which
are typically 1.75m wide and placed side by side, in a long line, for up to 400m in length
in order to support the roof of the coalface. Figure 2.38 shows the typical components
that make up a longwall shield.
Underground soft rock mining 111

Figure 2.37 The shearer, trailing power cable, AFC (a) and protective housing (Bretby) (b)
Photograph provided courtesy of Caterpillar. © Caterpillar Inc. All rights reserved.

Figure 2.38 T
 ypical components of a longwall shield (Mining Science & Technology, University of
Wollongong, 2020)
112 Underground soft rock mining

Figure 2.39 Typical frame type powered longwall support (Akande & Saliu, 2011)
1. hinge; 2. hydraulic control assembly; 3. leaf-spring thrusters; 4. centre base;
5. footplates with centering base; 6. shifting cylinder; 7. leg; 8. articulated canopy.

Figure 2.40 Powered chock longwall support


Photograph reproduced with kind permission from the Dowty Archive at Gloucestershire Archives,
collection reference D8347, from their website.

An individual chock can weigh 30–50 tonnes, extend to a maximum cutting height
of up to 8m, have a yield rating of 1000–1250 tonnes each and hydraulically advance
themselves typically 1m at a time.

• Frame support – a set of hydraulic jacks with common roof canopy and floor base
but designed with walking ability, which is completely independent of the face
conveyor or spill plates (Figure 2.39).
• Chock support – made of four to six hydraulic jacks connected by a common roof
canopy and floor base whose self-advancement depends on a common attachment
to a face conveyor or spill plates. An example is shown in Figure 2.40 (after Dowty).
Underground soft rock mining 113

• Shield support – consists of an inclined plate whose lower end is hinged to a hori-
zontal base plate sitting on the floor while the upper end is hinged to a horizontal
roof canopy in contact with the roof (Figure 2.41).
• Chock shield support – combines the advantages of the two-leg shields and the
four-leg chocks (Figure 2.42). The joint between the gob shield and the base is
mainly the lemniscates type, allowing the canopy to move straight up and down
and thus maximize support density and eliminate separate leg restoration.

2.5.4 Longwall entry (gate) design


The design of entries and the pillars between them are obviously very important.

2.5.4.1 Chain pillar design


Chain pillars can be either square, rectangular or in some cases diamond in shape.
They can also be in a single or double row configuration. In Australia, a single line of
rectangular chain pillars is the most common.
The intersections of gate roads and cut through are areas which require increased
roof support because of the large areas of exposed roof as well as increased roof
stresses at these locations. An increased frequency of intersections along a gate road
will increase the possibility of strata control difficulties.
Chain pillars carry minimal additional loading until the extraction of the adjacent
longwall panels. The strata load from above the extracted longwall panel is trans-
ferred from the longwall panel to the adjacent chain pillar(s). When the second adja-
cent longwall panel has been extracted, the chain pillars are subjected to extremely
high strata loads. In most situations, the pillars will collapse after the second longwall
panel has been extracted. This reduces the irregularity of subsidence at the surface.
The strength of a square or rectangular pillar can be determined from the following
formula:

kp 0.46
Pillar Strength ( MPa ) = (1)
h.66

Where k is a constant (typically 7.2 MPa), p is the pillar width (m) and h is the mining
height (m).

Figure 2.41 Longwall shield support and armoured face conveyor


Reproduced with kind permission from Komatsu Mining Corp.
114 Underground soft rock mining

Figure 2.42 Typical chock shield support (Akande & Saliu, 2011)
1. Front canopy 2. Gob shield 3. Base 4. Hydraulic ram 5. Spill plate 6. Hydraulic
leg/cylinder 7. Hydraulics and controls

The load that is transferred onto a pillar is calculated by a simple empirical for-
mula; the equation varies depending on the type of pillars and subsidence method
being used:
• Single line of square chain pillars
• Double line of square chain pillars
• Single line of rectangular chain pillars
• Double line of rectangular chain pillars

For rectangular chain pillars, if the width of one side of the pillar is greater than twice
the width of the other side, then the effective width method must be employed to calcu-
late the strength of the pillar.

• Single Line of Square Chain Pillars (Mining Science & Technology, University of
Wollongong, 2020)
Underground soft rock mining 115

Figure 2.43 Load on the chain pillars


Mining Science & Technology, University of Wollongong, 2020

The load transferred from the adjacent goaf areas is dependent upon the depth of
cover to satisfy subsidence compatibility (Figure 2.43).
For sub-critical subsidence conditions, the following expression is used to calculate
the load supported by a pillar:
For w/h < 2 tan φ:
 w 2 cotϕ 
Load on a Pillar = 9.81γ  ( p + w ) h − ( p + B) (2)
 4 

For w/h >= 2 tan φ:


For critical and super critical mining subsidence situations, the goaf load triangle will
reach the surface boundary; for this situation, the load on the pillar equation becomes:

Load on a Pillar = 9.81γ  ph + h 2 tanϕ ( p + B ) (3)

The respective average pillar stress for the above two conditions is given as:
For w/h < 2 tan φ:
 w 2 cotϕ  ( p + B )
Effective Pillar Stress (σ) = 9.81γ  ( p + w ) h − (4)
 4  p2

For w/h >= 2 tan φ:

( p + B)
Effective Pillar Stress (σ) = 9.81γ  ph + h 2 tanϕ (5)
p2

Where:

w = width of longwall face (m)


h = depth below surface (m)
γ = average density of overburden (kg/m3)
φ = average angle of shear (deg)
p = pillar width (m)
B = room (bord or heading) width (m)
116 Underground soft rock mining

Figure 2.44 Longwall roadway layout (Mining Science & Technology, University of Wollongong, 2020)

• Double Square Chain Pillars (Mining Science & Technology, University of Wollongong,
2020)
The only physical changes from a single line of square chain pillars is that the
effective width of the chain pillar support area becomes (2p + B) and the load is
borne by two pillars of a plan area 2p2 . The respective average pillar stresses for
this situation are:
For w/h < 2 tan φ:
 w 2 cotϕ  ( p + B )
Effective Pillar Stress (σ) = 9.81γ  ( 2p + B + w ) h − (6)
 4  2p2

For w/h >= 2 tan φ:

( p + B)
Effective Pillar Stress (σ) = 9.81γ  (2p + B)h + h 2 tanϕ (7)
2p2

Where:

w = width of longwall face (m)


γ = depth below surface (m)
g = average density of overburden (kg/m3)
φ = average angle of shear (degrees)
p = pillar width (m)
B = room (bord or heading) width (m). This is shown in Figure 2.44.

• Single Rectangular Chain Pillars (Mining Science & Technology, University of


Wollongong, 2020)

Rectangular chain pillars are employed extensively and frequently in preference to


square pillars in view of less junctions being formed during development. The average
pillar stress is given by:
Underground soft rock mining 117

For w/h < 2 tan φ:


 w 2 cotϕ  ( L + B )
Effective Pillar Stress (σ) = 9.81γ  ( p + w ) h − (8)
 4  p.L

For w/h >= 2 tan φ:

(L + B)
Effective Pillar Stress (σ) = 9.81γ  ph + h 2 tanϕ (9)
p.L

• Double Rectangular Chain Pillars (Mining Science & Technology, University of


Wollongong, 2020)

For w/h < 2 tan φ:


 w 2 cotϕ  ( L + B )
Effective Pillar Stress (σ) = 9.81γ  ( 2p + B + w ) h − (10)
 4  2p.L

For w/h >= 2 tan φ:

(L + B)
Effective Pillar Stress (σ) = 9.81γ  (2p + B)h + h 2 tanϕ (11)
2p.L

Where:
w = width of longwall face (m)
h = depth below surface (m)
γ = average density of overburden (kg/m3)
φ = average angle of shear (degrees)
p = pillar width (m)
L = pillar length (m)
B = room (bord or heading) width (m). This is shown in Figure 2.44.

2.5.5 The selection of design of longwall face supports


A complete shield support design consists of four main elements: amount of roof load
and lateral load to be supported, structural design and testing of the shield, determi-
nation of floor load distribution and service life of the supports. The following section
explains procedures for selection of longwall supports.

2.5.5.1 Determination of size of the support (Peng, 2020)


The following steps are typically required:
• Mining Height: Maximum and minimum working height of the support can be
determined using the following equations:

H max = H + ∆H-C-C r (1)

H min = H - ∆H-C-C r (2)


118 Underground soft rock mining

Where:

H = average seam thickness


ΔH = one-half of the maximum change in the thickness
C = roof convergence at the face area
Cr = height of the rock debris and asperities on the canopy and floor under-
neath the base plate

• Prop-Free Front: This refers to the distance from the face to the shield legs. It
usually ranges from 3.6m to 4.6m. The size of the prop-free front depends on the
width required to install cutting machine, face conveyor and space for crew travel.
• Ventilation Considerations: If it is a gassy mine, the methane gas released during
excavation needs to be diluted to safe levels. Supports with large lateral dimension
provide more airflow through the face, and therefore, released methane can be
diluted to safe levels.

2.5.5.2 Determination of roof loading on the support


Two factors are used to estimate height of the detached roof block: stratigraphic
sequence and bulking factor. The weight of the rock block that is supported by the
shield support is:

HBγ im L
Wf = (3)
Ko − 1

L = Lw + Lu + Lc + Lo (4)

Where:

H = mining height
B = centre-to-centre distance of the shield supports
γim = density of immediate roof
Lw is width of cut
Lu is unsupported distance
Lc is shield canopy length
Lo is rear overhang
Ko is bulking factor

2.5.5.3 Floor pressure under the base plate of a longwall shield (Figure 2.45)
(Taken from Peng, 2020)
• Rigid Base Plate: Depending on the location of the vertical resultant force under
the base plate, the pressure distribution in the floor can be simplified to triangle
shape, a trapezoidal or a rectangular shape.
If the pressure distribution is triangle (0<s<L/3), the magnitude of the pressure is:
Rv
pavg = × 2000 (4)
lw
Underground soft rock mining 119

Figure 2.45 Floor pressure distribution of a longwall shield base (Peng, 2020)

2Rv
p max = 2pavg = × 2000 (5)
lw

Where:

pavg = average floor pressure in psi


l = length of the floor pressure envelope in inch
pmax = the maximum floor pressure at the front end of the base plate in psi
w = width of the base plate

If s = L/3, the pressure distribution is still triangular, but l = L. The above equations
can be written as:

Rv
pavg = × 2000 (6)
Lw

2Rv
p max = 2pavg = × 2000 (7)
Lw
120 Underground soft rock mining

If L/3<s<L/2, the pressure distribution is trapezoidal, and the magnitudes of the pres-
sure are:

2Rv  3s-L 
p min =   × 2000
Lw  L  (8)

Rv  6  L  
p max = 2- s- × 2000
Lw  L  3   (9)

If s = L/2, the pressure distribution is rectangular, and the magnitude of the pres-
sure is:

Rv
p max = pavg = × 2000 (10)
Lw

• Elastic Base Plate: The maximum floor pressure for an elastic base plate is always
located directly under the vertical resultant force and the magnitude of the floor
pressure is nearly the same for the same resultant force, floor property and contact
width between the base plate and floor.
Based on the finite element analysis, a regression equation was developed to
determine the maximum floor pressure, which is as follows:

p max = 1233 + 213x1 + 2.72 x2 − 38x3 (11)

Where:

x1 is the Young’s Modulus of the floor material psi


x2 is the vertical resultant load in tonnes
x3 is the contact width between the plate and the floor along the faceline
direction in inches

2.5.6 Longwall subsidence


Longwall subsidence is the main issue associated with longwall mining, and the mech-
anisms involved must be understood. The subsidence not only causes surface topog-
raphy damage but also adversely affects the ground and surface water, which is often
even more serious.

2.5.6.1 General overview of surface subsidence associated with longwall coal mining
The maximum possible vertical subsidence/displacement (s) on surface above a long-
wall panel is created by the critical width of the longwall for any specific seam and
depth. If s is less than this maximum value, it is called subcritical, and if the maximum
value occurs over a volume as opposed to just a single point, it is called the supercrit-
ical (Figure 2.46).
Underground soft rock mining 121

Figure 2.46 Surface subsidence profiles above longwall mines (Gray et al., 1974)

Work by Jeran and Trevits (USBM RI 9552 of 1995) shows that the total longwall
subsidence is complete about a year after mining, and that subsidence above the centre
of the panel (where it is greatest) is 90% complete by the time the face has advanced
a distance equal to the overburden thickness. Most subsidence caused by longwall
mining occurs within a few days after coal extraction.
Most subsidence caused by longwall mining occurs within a few days after coal
extraction. The final compaction of the collapsed strata and surface subsidence is usu-
ally completed within one to two years after a longwall panel has been fully extracted.
The subsidence profiles are shown in Figure 2.46, which are explained in more detail
later.

2.5.6.2 Estimating longwall surface subsidence


• Overburden zones (affected by the longwall mining-induced subsidence)
122 Underground soft rock mining

Figure 2.47 The fractured zones above a longwall panel (Peng, 2008)

Depending mainly on the depth, there are typically four overburden zones (Figure 2.47)
that are distinguished by the level of the severity of the strata disturbance (Peng, 2008).
They consist of (where h is the seam height mined):

• Caving Zone 1
• Fracture Zone (30–50H) – Zone 2
• Bending/Continuous Deformation Zone (>60h) – Zone 3
• Surface Subsidence Zone – Zone 4

2.5.6.3 The surface subsidence zone


The following are general “rules of thumb” and issues caused by surface subsidence:

• The maximum subsidence on surface is typically around 60–70% of the seam


thickness.
• In flat-lying areas, this can cause localized “ponding” (water to collect in the
depression formed).
• The surface is also stretched, compressed and tilted (depending on the part of the
subsidence trough (refer to Figure 2.48).

The surface effects are directly proportional to the coal seam mined thickness and
inversely proportional to depth.
In the surface extension zone, the following can form in the surface soil layer close
to the perimeter of the longwall panel:

• Cracks can form up to 15m deep as the surface layer is stretched.


• Permeability increases (due to these cracks and fractures) often with major adverse
effects on surface water, aquifers and the water table.

2.5.6.4 Some relevant definitions (Peng, 2020)


The Subsidence (S) refers to vertical displacement of a point, but subsidence of the
ground actually includes both vertical and horizontal displacements. These horizontal
Underground soft rock mining 123

Figure 2.48 S ubsidence above a longwall coal mine showing the internal angle-of-draw (the caving
angle) (Michaud, 2018).

displacements can in some cases be greater and more damaging than the vertical sub-
sidence, where the subsidence is small.
The Angle-of-Draw or Limit Angle (A) is the outward angle between the normal to
the seam at the panel edge and a line connecting the panel edge and the point on the
surface where the observed subsidence is zero (Figure 2.49). All of the surface subsid-
ence occurs within the angle-of-draw.
Note that the maximum surface subsidence (generally 50–60% of the coal seam mined
height) will only occur at a width equal or greater than the critical longwall width.
The Maximum Possible Subsidence (Smax) is the maximum vertical surface displace-
ment for a panel that is wide enough so that the edges do not limit the subsidence in the
centre of the panel (a critical or supercritical panel) as shown in Figure 2.49.
The Internal Angle-of-Draw (A’) is the internal angle between the normal to the
seam at the panel edge and a line connecting the panel edge and the point on the
124 Underground soft rock mining

Figure 2.49 The limit angle or angle-of-draw (Anderson et al., 2006)

surface interior to the panel where the observed subsidence is no longer affected by
the panel edge. For a wide panel, the Internal Angle-of-Draw delineates the area of
maximum possible subsidence. Often the Internal Angle-of-Draw is considered to be
equal to the “external” Angle-of-Draw, although there is no reason these angles have
to be exactly equal (Figure 2.49).
The Inflection Point on the subsidence curve is the point where the downward curve
at the edge of the panel changes to the upward curve going to the middle of the panel.
The Inflection Point is where the surface slope is maximum and the strain changes
from tension to compression. The Inflection Point is generally considered to be the
point of half subsidence (S/2) and to occur slightly in from the panel edge (Figure 2.49).
The Subsidence Element (a) is the percentage of the seam thickness (h) that shows
up as the maximum possible vertical displacement (Smax) on the surface.

Smax = a * h

Figure 2.50 shows the maximum subsidence and the tension and compression zones
above a longwall panel (Peng, 2008). Figure 2.51 shows the inflection points near the
edge of a longwall panel (Hustrulid, 1982).
The Slope (Σ) is the change in subsidence over the change in distance (dS/dx). It is
the derivative of the subsidence.

dS S2 − S1
Σ= =
dx x2 − x1

Where:

Σ is the surface slope


S1 the subsidence at Point 1
Underground soft rock mining 125

Figure 2.50 Maximum possible subsidence (Smax) (Peng, 2008)

Figure 2.51 The longwall subsidence inflection point (Hustrulid, 1982)

S2 the subsidence at Point 2


x1 the location at Point 1
x2 the location at Point 2

The Differential Slope (θ) is equal to the change in the slope over the change in
distance.

dΣ  (S3 − S2 ) (S2 − S1 )   2 
θ = = − *  
dx  x3 − x2 x2 − x1   x3 − x1 
126 Underground soft rock mining

Where:
θ is the differential slope
Σ is the surface slope
S1 the subsidence at Point 1
S2 the subsidence at Point 2
S3 the subsidence at Point 3
x1 the location at Point 1
x2 the location at Point 2
x3 the location at Point 3
For a constant distance (dx) between subsidence measurement points, the Differential
Slope (θ) is equal to:

dΣ  (S3 − S2 ) (S2 − S1 )  2
θ= = −  *
dx  x 3 − x 2 x 2 − x1  x 3 − x1
 (S − S2 ) (S2 − S1 ) 
=  3 − * 2
 dx dx  2 * dx
 
 (S − 2S2 + S1 ) 
= 3 
 dx 2

The Radius of Curvature (r) is the inverse of the Differential Slope (θ).

1  dx 2 

r= = 
θ  (S3 − 2S2 + S1 ) 

Where:
r is the radius of curvature
θ is the differential slope
S1 the subsidence at Point 1
S2 the subsidence at Point 2
S3 the subsidence at Point 3
dx is the distance between points.
The Strain (ε) is caused by bending and differential horizontal movements in the strata
above the longwall panel. Strain is reliably measured from monitored and calculating
the horizontal change in length of a section of a subsidence profile and dividing this
by the initial horizontal length of that section. If the section has been extended, the
ground is in tension and the change in length and the resulting strain are generally
taken as positive (by convention). If the section has been shortened, the ground is in
compression and the change in length and the resulting strain are negative. The sign
convention does not matter as long as it is consistently applied.
The strain is proportional to the change in the slope over the change in distance.

dΣ (S3 − 2S2 + S1 )
ε = * 0.024 = * 0.024
dx dx 2
Underground soft rock mining 127

Figure 2.52 The effects of an advancing longwall face (Ingram, 1989)

Where:
ε is the surface strain
Σ is the surface slope
S1 the subsidence at Point 1
S2 the subsidence at Point 2
S3 the subsidence at Point 3
dx is the distance between points
The Tilt is calculated as the change in subsidence between two points divided by the
distance between those points. Tilt is, therefore, the first derivative of the subsidence
profile. The sign of tilt (positive or negative) is not important, but the convention usu-
ally adopted is for a positive tilt to indicate the ground increasing in subsidence in the
direction of measurement. The maximum tilt, or the steepest portion of the subsid-
ence profile, occurs at the point of inflection (contraflection) in the subsidence trough,
where the subsidence is roughly equal to one half of the maximum subsidence. Tilt is
usually expressed in millimetres per metre (see Figure 2.52).
The most significant impacts caused by a longwall on surface infrastructure are
experienced during the development and movement of the subsidence trough, when
maximum ground movements normally occur. As the subsidence wave approaches a
point on the surface, the ground starts to settle, is displaced horizontally towards the
mined void and is subjected to tensile strains, which build from zero to a maximum
over the length of the convex curvature, as shown in Figure 2.53.
The amount of ground subsidence that occurs is dependent mainly upon the width
of the underground longwall panel and the depth of mining. The degree of subsidence
that occurs is classified into three categories:
• Sub-critical Width
• Critical Width
• Super-critical Width
Figure 2.54 gives the relationship between subsidence and tilt based on measurements
(Hunt, 1979).
128 Underground soft rock mining

Figure 2.53 The development of the subsidence trough above an advancing longwall panel
Source: Mine Subsidence Engineering Consultants (2007)

Figure 2.54 The relationship between subsidence and surface tilt (Hunt, 1979)
Underground soft rock mining 129

Figure 2.55 Sub-critical subsidence (Lee & Abel, 1983)

2.5.6.5 Sub-critical panel


The term Sub-critical refers to a panel where the width (W) is so narrow in relation
to the depth (H) that the lines denoting the internal angle-of-draw/caving angle (A’)
cross before they reach the surface (Figure 2.55). The maximum possible subsidence
is therefore never obtained.

W
< 2 tan A'
H

2.5.6.6 Critical panel


As a longwall panel width is increased, the overlying strata are less able to arch over
the goaf/gob and a limiting panel width is reached where no support is available and
maximum subsidence occurs. This limiting panel width is referred to as the critical
width and is usually assumed to be 1.4 times the depth of cover. It does, however,
depend upon the nature of the strata and other factors (especially mining height). A
critical panel (Figure 2.56) has exactly the required ratio of width (W) to depth (H) so

Figure 2.56 Critical subsidence (Lee & Abel, 1983)


130 Underground soft rock mining

Figure 2.57 Super-critical subsidence (Lee & Abel, 1983)

that the lines denoting the internal angle-of-draw just meet at the surface. The max-
imum possible subsidence is obtained at exactly one point in the middle of the panel.

W
= 2 tan A′
H

2.5.6.7 Super-critical panel


A Super-critical panel (Figure 2.57) is so wide in relation to the depth (H) that the lines
denoting the internal angle-of-draw do not cross or meet before they reach the surface.
The maximum possible subsidence is obtained at all points between the surface inter-
sections of the internal angle-of-draw lines.

W
> 2 tan A′
H

2.5.6.8 Sub-surface displacements and strains


Figure 2.58 shows the areas of extension (tension) and compression above a typical
longwall panel.

2.6 Specialized mining methods

2.6.1 Coal mining with backfill


The goal of backfilling on coal mines is to protect the ground and surface water and
limit surface subsidence.
Subsidence refers normally to a surface point sinking to a lower level. It usually
refers to a vertical displacement, but in longwall mining, a horizontal movement also
occurs close to the edges of the longwall panel.
Subsidence is a process that occurs in response to voids being created by extracting
solids or liquids from beneath the Earth’s surface. This can be a natural or man-made
process. In mining, it is controlled by many factors, including:

• Mining method used


• Depth of and extent of the mining
Underground soft rock mining 131

Figure 2.58 H
 orizontal deformation of a strata continuum over longwall mining in a bedded deposit
(Kratzsch, 1983)

• Effective thickness of the deposit


• The topography
• In-situ properties of the rock mass above the deposit
• The geology and hydrology

The impacts of subsidence are potentially severe in terms of damage to surface util-
ity lines and structures, changes in surface-water and ground-water conditions and
effects on vegetation and animals.
Most coal mines currently use surface waste disposal, which can contaminate
the surface and ground water and creates the real potential for a catastrophic fail-
ure of the waste facility. Coal mining is one of the few mined materials where the
waste produced can all be placed back underground (similar to limestone, halite
mining, etc.). This has the obvious environmental benefits if done correctly as
it improves the surface environment and can drastically reduce mining-induced
subsidence.
It must be noted that subsidence cannot be totally eliminated using backfill but it
can be greatly reduced so that it has no significant effect on water features or surface
infrastructure. Backfill can also increase costs significantly so its use must be justi-
fied. This justification usually involves the protection of the environment (water and
infrastructure).

2.6.1.1 Backfilling a room and pillar coal mine


In room and pillar coal mining, a regularly spaced array pillars of coal are left to sup-
port the overburden. Openings (rooms or bords) are driven orthogonally and at reg-
ular intervals to leave typically square pillars for the overburden support. The stress
regime (especially the horizontal stress which can be higher than the vertical stress)
132 Underground soft rock mining

can create potential mining stability challenges. The stability of the overall system can
be adversely influenced by aspects such as:

• Incorrect design
• Undersize pillars (especially with corners removed due to the turning radius of
the continuous miner)
• Oversized rooms
• Long term weathering.
• Weak floor or roof

The coal pillars left to support the overlying strata can therefore deteriorate over time
due to weathering caused by humidity and water ingress especially after mining ends.
The progressive pillar deterioration reduces its strength, effective size and shape, thus
negatively affecting the pillar integrity and stability. This can cause subsidence issues
in the long term (Spearing et al., 2013).
Where secondary pillar removal is undertaken, backfill can greatly assist in main-
taining the stability, but effective placement can be quite involved and costly.
Sinkhole subsidence is common in areas overlying shallow room and pillar mines.
Sinkholes occur from the collapse of the mine roof (especially at an intersection) into a
mine opening, resulting in caving of the overlying strata and an abrupt and steep depres-
sion forming at the ground surface. The majority of sinkholes usually develop where the
amount of cover (vertical distance between the coal seam and the surface) is less than about
50m. This type of subsidence is generally localized in extent, affecting a relatively small
area on surface. However, structures and surface features affected by sinkhole subsidence
tend to require extensive and costly damages. Sinkhole subsidence has been responsible
for extensive damage to numerous homes and property throughout the years in the US.
Sinkholes are typically associated with old and abandoned shallow mine workings.
Based on current regulations in the US, for example, the Mine Safety and Health
Administration (MSHA) will not authorize underground mining beneath structures
where the depth of overburden is less than about 30m, unless the subsidence con-
trol plan demonstrates that proposed mine workings will be stable and that overlying
structures will not suffer irreparable damage over time. An example of the subsidence
is shown in Figure 2.59.

Figure 2.59 C
 ommon form of sinkhole and trough subsidence in room and pillar mines (Whittaker
& Reddish, 1989)
Underground soft rock mining 133

Figure 2.60 Trough subsidence (Whittaker & Reddish, 1989)

Subsidence troughs (Figure 2.60) induced by room and pillar mining can occur
above active or abandoned mines. The resultant surface impacts and damages can be
similar; however, the mechanisms that trigger the subsidence are different. In aban-
doned mines, troughs usually occur when the overburden sags downward due to the
failure of numerous remnant mine pillars, or by punching of the pillars into weak
mine floor or roof. It is difficult, if not impossible, to predict if or when failure in an
abandoned mine might occur, since abandoned mines may collapse many decades
after the mining is completed if the mine workings were not designed to provide long-
term support. Water inflow after mining ceases creates ideal weathering conditions for
the coal pillars, the roof and the floor. Figure 2.61 shows actual trough subsidence in
farmland caused by coal mining.
Backfilling can provide long-term stability to room and pillar panels by produc-
ing a compressive (confining) pressure to the pillars that could increase over time
if the pillar started dilating (failing and bulking). This would drastically reduce the
chance of instability. Spearing et al. (2013a) demonstrated that high-density back-
fill in room and pillar coal mines in the US was technically feasible, and although
not currently cost effective, it is likely to become cost effective in the medium
term.

Figure 2.61 Trough and sag subsidence above an abandoned coal mine in New South Wales, Australia
Photograph reproduced with kind permission from the Subsidence Authority, NSW (www.subsidenceadvisory.
nsw.gov.au)
134 Underground soft rock mining

Kostecki and Spearing (2015) found that a 10–40% increase in pillar strength was
observed when the pillars were backfilled to a height of 25–75% respectively.
The key factor associated with distribution complexity and cost in a room and
pillar panel is the beach angle that can be achieved with quality backfill as this
directly influences the number of discharge points required. The number of dis-
charge points then dictates the piping required to adequately backfill a panel, and
this pipeline cannot be recovered. The beach angle is the angle at which backfill
material naturally comes to rest after placement. A coal mine in Australia has
produced a high density backfill material for a longwall application with a beach
angle between 4° and 5°.

2.6.1.2 Backfilling a longwall coal mine


As mentioned earlier, subsidence from longwalling mining is one of the biggest
adverse problems of longwalling mining and the profiles are shown again on
Figure 2.62.
Backfilling is used on longwalls to limit subsidence and protect the surface and
ground water.
About a third of Chinese coal reserves are under major water features, roads, rail-
ways and cities.

Figure 2.62 S urface subsidence profiles above longwall panels of different widths (Gray et al.,
1974)
Underground soft rock mining 135

The School of Mines at the China University of Mining & Technology (CUMT)
developed “Three Down Categories” that refer to coal reserves below sensitive
infrastructure:

• First Down Category: under buildings coal accounts for more than 60% of the
whole “three under” coal amount.
• Second Down Category: under water (including pressure water above) coal
accounts for about 28%.
• Third Down Category: under railway/road under pressure coal accounts for
about 12%.

To make these reserves mineable, quality backfilling to reduce surface subsidence by


+85% has been researched by the School of Mines at CUMT and commercialized
successfully multiple times.
According to Zhang et al. (2019), while supporting socio-economic development,
the constant large-scale mining of coal resources causes many problems, such as the
ecological destruction of mining areas and accelerated depletion of resources. On the
one hand, coal mines produce a relatively large amount of solid wastes (gangue) whose
discharge accounts for about 10–15% of coal production.
The general handling method is to discharge and accumulate the gangue on the
surface in gangue dumps. Gangue dumps not only occupy land resources but also
pollute the air and underground water and even result in potential disasters such as
landslides and explosions. On the other hand, high-intensity longwall mining of coal
often induces a series of geological and environmental disasters.
By using the longwall caving method to handle goafs, the roof loses the support
of the lower coal seams, subsidence and breakage occur under the self-weight of the
strata and the effect of overburden pressure constantly develop to overly the strata,
causing surface subsidence, thus resulting in problems such as soil and water loss, land
subsidence, vegetation dieback and building collapse.
To solve the problems caused by the high-intensity mining of coal, Zhang et al.
(2019) in China proposed a coal mining method with backfilling of goafs. Backfill
materials prepared with solid wastes, such as gangue and fly ash, are directly back-
filled into goafs in each backfill mining panel to replace coal resources and control
strata movement and surface subsidence, thus achieving the goal of protecting the
environment above mining areas.
• Hydraulic Backfilling
Backfilling behind the advancing deep longwall face has been commonly used in
Germany (the Pruessag pumped system) and Poland for decades using mainly hydrau-
lic sand filling (Palarski, 1989). Paste backfilling performs better for subsidence con-
trol because it has high solids density and no post-filling shrinkage due to subsequent
water drainage (loss) after placement.
Backfill placing fly ash slurry behind a longwall has been successfully used in China
too as shown on Figure 2.63 (Jiang et al., 2017).
• Solid Backfill Mining (SBM)
Solid backfill technology can virtually eliminate surface subsidence above long-
walls. This has allowed coal seams to be mined under high-speed railway lines,
136 Underground soft rock mining

Figure 2.63 Placing fly ash behind a longwall face (Jiang et al., 2017)

buildings and water, which would have been impossible without SBM (Zhang
et al., 2019).
In essence, waste is pre-processed to ensure low porosity, then brought under-
ground dry (or pre-damped) and placed behind the advancing longwall face by an
additional automated conveyor. The waste is dropped sequentially through each
shield and a hydraulic ram mounted on the rear of the shield helps compact it into
a free-standing wall which provides support to the immediate roof that would
otherwise goaf extensively. A schematic solid backfill mining system is shown on
Figure 2.64.
To date, over 100 longwalls panels have been mined in 40 different coal mines
throughout China, and it is now an accepted method when mining below sensitive
areas. This method has therefore been used in situ numerous times and has been
accepted by the Chinese coal producers as a viable method above sensitive areas that
would otherwise remain unmined.

Figure 2.64 Schematic layout of the solid backfilling system (Zhang et al., 2019)
Underground soft rock mining 137

Figure 2.65 Sketch of the modified longwall shield used for solid backfilling (Zhang et al., 2019)

Modified longwall shields (see Figures 2.65 and 2.66) are used with an additional con-
veyor on the back of the shields with gates to drop the solid backfill in the correct place in
sequence and a hydraulic rammer to set the fill tightly against the roof (Feng et al., 2017).
The effect of solid backfill mining is shown diagrammatically in Figure 2.67. The
backfill significantly reduces the gravity-induced goafing (caving).
Figure 2.68 shows the dry solid backfill being placed behind the shields and com-
pacted hydraulically.
The performance of uncemented backfill depends mainly on the porosity and the
mechanical properties of the fill material. Figure 2.69 ( Huang et al., 2020) shows tests
of some locally available waste products for a specific longwall mine in China.

Figure 2.66 An actual modified longwall shield


Reproduced with kind permission from School of Mines, China University of Mining & Technology
138 Underground soft rock mining

Figure 2.67 A
 comparison between longwalling without backfill and longwalling with solid backfilling
(Zhang et al., 2019)

Figure 2.68 Placing the backfill behind the shields (Zhang et al., 2019)
Photograph reproduced with kind permission from the School of Mines, China University of Mining & Technology

Figure 2.69 T
 he stress-versus-strain graphs for various locally available waste products for backfilling
(Huang et al., 2019)
Underground soft rock mining 139

Figure 2.70 Material transportation schematic (Zhang et al., 2019)

Typical transportation routes for the coal coming out of the mine and the backfill
waste coming back into the mine are shown in Figure 2.70.
The preparation and transport layout for an uncemented waste fill system is shown
in Figure 2.71.

2.6.2 Top coaling


Ultra-high seams (or high height seams) are loosely categorized as being over 6m.
Mining these seams is more complicated, and top coaling is commonly used.

Figure 2.71 Uncemented backfill plant layout (Zhang et al., 2019)


140 Underground soft rock mining

Figure 2.72 Top coaling using a single AFC (Solymos, 1983)

Initially in the literature from the 1960s, top coaling was called soutirage (a French
name) but changed in the 1990s to top coaling (Proust, 1979).
Top coaling has been used successfully in coal seams up to 30m in height (Zhou
et al., 2020).
Top coaling or top caving is used when the coal seam height is too large to permit
a safe and stable single cut extraction operation. Taking a single high cut in long-
walls often requires much more costly equipment which is much heavier and can cause
handling and operational issues. The main advantages of top coaling are relatively
low costs and high recovery. It is typically considered for coal seams that are more
than 4.5m thick, but more common above 6m. The equipment is basically the same as
in a conventional single cut longwall except the shields are modified and fitted with
an extra chute (if a single AFC is used as shown in Figure 2.72) or an extra chute with
an additional AFC as shown in Figure 2.73.

Figure 2.73 Top coaling using two AFCs (Austar Mine, Yancoal, Australia, 2006)
Underground soft rock mining 141

Having two AFCs has the major advantage of much higher production. The key,
however, is to coordinate the rate of shearing and hence advance with the time neces-
sary for the top coal to goaf and be effectively collected on the AFC. Automation is
critical to help achieve an optimum production rate. The opening and closing of the
slides (gates) to let the goafed top coal fall under gravity on the AFC is very important
and needs to be effectively controlled. Typically, the slides are opened in a logical
sequence from one shield to the next, removing the top coal and closing once waste
rock is detected (above the top coal roof).
The potential advantages of top coaling are:

• Good recovery but at a lower effective mining height resulting in improved safety
• Improved active longwall face stability
• Reduced costs
• Easier and less costly longwall gate (entry) development

The main disadvantages are:

• Possible increased risk of spontaneous combustion from broken coal left in the
gob (goaf)
• Lower extraction and productivity than cutting the seam in a single full height
pass

The method was conceived in Europe, but China has led the development of top coal-
ing now, and they have more than 200 top coal longwall panels. The system has also
been introduced in Australia and Turkey.

2.6.3 Ultra-high seams with single cut longwalling


To reduce the spontaneous combustion hazard and increase extraction, ultra-high
seams with single cut longwalling was developed in China in 2009 and introduced
initially in the Bulianta Mine. Longwall shields 7 m high were developed and success-
fully used for the application. Current shields can now handle coal seams of almost
9m, as shown in Figure 2.74.

Figure 2.74 Example of some +8m high longwall shields


Reproduced with kind permission from ShangWan Coal Mine, China.
142 Underground soft rock mining

The main challenges using this method are:


• The longwall can goaf suddenly and not evenly, which at the increased height can
cause problems.
• A localized or major face collapse can occur, caused typically by adverse geology,
gas outburst or high stresses.

2.6.4 Multi-cut coal mining


High coal seams can be mined as follows:
• In a single full cut
• By top-coaling
• By mining slicing
The selection depends mainly on the stress tensor and the coal seam properties.
Multi-cut coal mining of wide coal seams with backfill has been used successfully
in Poland (Spearing, 1994). Typically, a top longwall cut is taken and backfilled with
successive follow-behind backfill cuts as shown in Figure 2.75. The individual cuts are
usually between 2.0m and 3.5m.
According to Palarski (2010), multi-cut coal mining can, however, be undertaken
in various ways depending mainly on the coal seam dip (see Figure 2.76). The most
common methods are:
• Taking the slices parallel to the roof and floor.
• Generally mining up-dip to facilitate the backfill placement.
• The order of slicing can be descending or ascending.
• The coal seam slices can be undertaken concurrently or independently.
Palarski (2010) states the following for ascending or descending slices:

2.6.4.1 Descending slices


This method is used in thick seams that are relatively weak. The first slice is cut against
the coal seam roof and is backfilled concurrently with a cemented fill material. The
second and subsequent slices are mined underneath an artificial (backfilling roof) in
a way similar to hard rock undercut and fill.

Figure 2.75 Multi-cut longwall coal mining (Palarski, 2010)


Underground soft rock mining 143

Figure 2.76 Multi-cut coal longwall options (Palarski, 2010)

The advantages of the descending method are:

• The coal longwall shields rest on the coal, which makes a relatively firm support base.
• The fire and methane emission hazards are minimized.
• The backfill can use a large amount of very fine backfill material.

The disadvantages of the method are:

• The time between the extraction of successive slots is relatively long as the backfill
in the slot above must first gain adequate strength.
• Shield or chock supports must have large and wide canopies.
• Often 20–50 cm of coal must be left behind after the first slot to reduce the backfill
strength requirement and the costs.

2.6.4.2 Ascending slices


Two variations of this method are used in Polish coal mining:

• Non-concurrent mining of each slice


• Concurrent mining of slices

For non-simultaneous ascending slicing with backfill:

• Development workings: In the first floor slice, development workings are carried
out in a way similar to mining any medium thick coal seam. In the next slice, the
workings are carried out in a similar way but shifting the entries (gates) horizon-
tally by at least the width of a gate.
• Longwall faces in upper slices are different to those used to extract a medium coal
seam conventionally.
144 Underground soft rock mining

The shields or chocks have large support bases to reduce the chance of them sink-
ing into the backfill floor after the first slice has been extracted. Backfill barricades
must be dug into the previously placed fill, and drainage arrangements must be made
to avoid backfill wash-out or liquefaction. Spontaneous combustion of coal is more
of a hazard and must be monitored very carefully. Tight backfilling is important and
entries that are not needed anymore must be immediately filled and/or sealed.

2.7 Support design and application

2.7.1 Coal entry and room support design

2.7.1.1 Support challenges and requirements


The support challenges in weak rock are significant, and the following should be
considered:
• These excavations are difficult to analyze as failure is typically progressive and
weathering can also be a significant factor.
• Even relatively low stresses can cause local failures.
• Typical roof and floor rock types include mudstones, siltstones, tuffs and some
especially laminated shales, which are all relatively weak and prone to weathering.
• Deformation monitoring is essential, and arches (or steel sets), shotcrete and cast
concrete are commonly used. Shotcrete is, however, not too effective due to the
weak internal shear strength of particularly the coal.
Support in coal mines is generally classified as either primary support, secondary support
or supplemental support. Primary support is rock bolts that are installed in the devel-
opment cycle. Secondary support typically consists of cable bolts, trusses and/or stand-
ing support and is installed sometime after the initial development cycle, commonly in
intersections. Supplemental support is support installed by exception over and above
the primary and secondary support when conditions are worse than those designed
for or expected, due to, for example, weathering or for rehabilitation of an excavation.
Supplemental support is routinely installed after a longwall panel is completed and the
headgate is converted into the tailgate for the next adjacent longwall panel.
A successful design involves:

• Not only a safe solution although this is the key; the solution adopted must also
meet the functional requirements of the excavation and also represent a cost-ef-
fective solution.
• Engineering judgment based on available information is critical, and the design
must be revisited and modified as more real performance data becomes available.

There are three main support methods that account for how rock anchors work in
supporting the rock around excavations (especially the roof). These are:

• Key block (Figure 2.77)


• Beam building (Figure 2.78)
• Suspension (Figure 2.79)

They are important and simple concepts that are useful.


Underground soft rock mining 145

Figure 2.77 Key block support (Panek, 1956)

Figure 2.78 Beam building (Panek, 1956)

Figure 2.79 Suspension (Panek, 1956)


146 Underground soft rock mining

According to Spearing et al. (2013), the design of support must also take into account
the “human factor”. This means that the support system must make allowance for the
following possibilities:

• The tunnel may not always be adequately “made safe” (i.e., not all the loose rock
may be removed after each blast).
• The development crews may not always be able to identify potentially hazardous
conditions.
• Unplanned over-break.
• The support may not be correctly installed.

2.7.1.2 Support design methodologies (Spearing et al, 2013b)


There are several methods that can be used and the selection should be made based on
the complexity and uniqueness of the specific conditions.

EXPERIENTIAL METHOD

This design approach uses experience gained from other similar situations and conditions.
Experiential design for coal mine support is a very useful design tool especially for
brownfields operations. It is generally fair to assume that if a certain support system
worked under similar conditions before, it will still work.
The main problems are in the assessment of “similar conditions” and that “older
support types and methods” may be less effective and/or economic now.

EMPIRICAL METHOD

Many mine-related designs are based on empirically determined relationships. One


support design system that is empirically based is the CMRR developed at the US
Bureau of Mines (Mark, 1999). This was developed because other popular systems
seemed better suited to hard rock applications.The data required for the CMRR can
be determined either from underground exposures such as roof falls and overcasts or
from exploratory drill cores. In either case, the main parameters involved are:

• The uniaxial compressive strength (UCS) of the intact rock


• The intensity (spacing and persistence) of bedding and other discontinuities
• The shear strength (cohesion and roughness) of bedding and other discontinuities
• The moisture sensitivity of the different strata involved particularly in the imme-
diate roof
• The presence of a stronger rock bed in the bolted interval

Other secondary factors include the number of layers, the presence of groundwater
and surcharge from overlying weak beds.
The in-situ properties that have the greatest effect on intersection stability can be
characterized using the CMRR method. These parameters combined with the magni-
tude and orientation relative to the entry of the maximum horizontal stress (σhmax) can
provide an empirical basis for anticipating stability problems. The CMRR evaluates
defects, discontinuities and strength of the immediate roof strata. The rock samples
Underground soft rock mining 147

are assigned a number between 0 and 100 that can be used in engineering calculations
to help in pillar design, intersection sizing and other applications. The simplest way to
effectively utilize CMRR is to use the software of the same name, available from the
National Institute of Occupational Safety and Health (NIOSH) in the US. Typically,
roofs with CMRR<25 tend to fail very soon after mining, so the practical working
range is 25–100. Case studies confirm that a higher CMRR is related to lower num-
ber of roof falls. Even a difference of 10 can affect roof stability significantly. For
example, a mine in West Virginia had a weak shale (CMRR = 40) and a rock fall
rate that was 3.4 times more than a nearby mine with a stronger shale (CMRR = 50)
(Mark, 1999).
Other factors that also have an influence on the intersection stability include the
bolt length, stiffness, spacing, strength, orientation, installation quality and the tim-
ing of the support installation after the entry is developed.
Another commonly used empirically based system is the Hoek-Brown criterion
(Hoek & Brown, 1980). The following is from an unpublished note by Pillalamarry
and Spearing. Hoek-Brown’s empirical failure criteria has been widely adopted for
modelling the strengths of a wide range of rock types. The criterion was derived from
the results of research into the brittle failure of intact rock by Hoek and studies on
joint rock mass behaviour by Brown. However, the Hoek-Brown criterion also has its
limitations as do all empirically developed relationships. In particular, the controlling
parameters and the empirical expressions that link these parameters to the observed
mineralogy and rock structure can provide challenges. Despite these potential limi-
tations, the Hoek-Brown criterion has been widely accepted by the international rock
mechanics community. This criterion has been subject to modifications over the years.
The relationship is as follows:

 a
′ ′  σ3′ 
σ1 = σ3 + σci  mb + s (1)
 σci 
 
 GSI − 100 
mb = mi exp   (2)
 28 

Where:

σ1 is the major principal stress at failure


σ3 is the minor principal stress at failure
σci is the UCS of the intact rock material in the specimen
s and a are constants which depend upon the properties of the rock mass
mb is rock mass strength characteristics
m i is the value of the Hoek-Brown constant for the rock mass
GSI is Geological Strength Index (GSI)

The Hoek-Brown criterion parameters are determined as follows.


• UCS and Hoek-Brown constant (σ ci & m i): UCS and Hoek-Brown constant
are determined from results of the triaxial test conducted on intact rock.
The test should be carried out over a confining stress range from zero to one
half of the UCS. Five or more tests should be conducted to determine σci
148 Underground soft rock mining

and m i more accurately, and the values are determined using the following
equations:

∑ y  ∑xy − ( ΣxΣy/n )  ∑x
σ2ci = −  (3)
n  ∑x 2 − ( Σx )2 /n  n

1  ∑xy − ( ΣxΣy/n ) 
mi =   (4)
 
σci  ∑x 2 − ( Σx )2 /n 

Accuracy of the σci and m are determined using coefficient determination (r),
which is given as:

 ∑xy − ( ΣxΣy/n ) 2
r2 =  
(5)
( )  ( )
 Σx 2 − ( Σx )2 /n   Σy2 − ( Σy )2 /n 


Where:

X = σ′3 and y = ( σ′1 − σ′3 )2 (6)

Hoek et al. (1995) introduced the GSI to provide a system for estimating the reduc-
tion in strength for different geological conditions. The strength of jointed rock
mass depends upon the properties of the intact rock particles and also upon the
freedom of these particles to rotate and slide under different stress conditions. This
freedom is controlled by geometric shape of rock particles and the condition of sur-
face separating these pieces. The system is shown in Figure 2.80, from which GSI
value can be determined.
Once the GSI has been estimated, the parameters s and a are determined using
the following relationships:
For GSI>25, i.e., rock masses of good reasonable quality:
 GIS − 100 
s = exp   (7)
 9 
a = 0.5 (8)

For GSI<25, i.e., rock masses of very poor quality:

s=0 (9)
GSI
a = .65 − (10)
200

Various important parameters can influence the results:


• Effect of Water: There is compelling evidence that the strength of rock decreases
with pore water pressure in rocks. Hence, before using the Hoek-Brown criterion,
it is necessary to determine the pore water pressure and ultimately effective stress.
Underground soft rock mining 149

Figure 2.80 Geological Strength Index (GSI) (Hoek et al., 1995)

The stress parameters in equation (1) should be taken as effective stress rather
than absolute stress. The effective stress for a porous rock mass is determined
using following relation:

Effective stress ( σ′ ) = σ − u (11)


150 Underground soft rock mining

Where:

σ′ is the effective stress


σ is the stresses in rock mass
u is the pore water pressure

The pore water pressure is determined using piezometers or from manually con-
structed or numerically generated flow nets.
• Effect of the Sample Size: Generally, there is a significant reduction in the strength
with increasing sample size (due to rock imperfections and geological features).
Hoek-Brown suggested that the UCS σcd of a rock specimen with a diameter of “s”
mm is related to the UCS σc50 of a 50mm diameter sample by the following equation:

 50  0.18
σcd = σc 50   (12)
 d 

All empirical approaches have limitations, and Figure 2.81 illustrates the situations
when Hoek-Brown failure criterion should and should not be applied. The situations
where the Hoek-Brown criterion can and cannot be applied are as follows:

• When the volume under consideration is small enough that it does not contain
any structural discontinuities, Hoek-Brown failure equation can be applied, using
the m and s values for the intact rock. This condition would apply to small-scale

Figure 2.81 Application of Hoek-Brown criterion to different scales of rock mass (Hoek et al., 1995)
Underground soft rock mining 151

specimens which have been extracted for laboratory testing or to the analysis of
concentrated forces such as those which may be exerted by an individual pick on
a tunnel boring machine cutter.
• When the volume of rock being considered is such that only a few structural discon-
tinuities are contained in this volume, the Hoek-Brown failure criterion should not
be used. The behaviour of this rock is likely to be highly anisotropic, and the Hoek-
Brown failure equation is only applicable to isotropic rock and will give erroneous
results.
• When the volume under consideration contains four or more closely spaced dis-
continuity sets and none of these discontinuity sets is weaker than any of the
other, then the Hoek-Brown criterion can be used. If one of the discontinuities
is very weak as compared to the other, the rock mass should be treated as aniso-
tropic and Hoek-Brown criterion should not be applied unless allowance is made
for this anisotropy.

It is important that a support designer finds a system that seems useful for the specific
application but the key is to monitor, analyze and revisit the design as more informa-
tion becomes available. Updating the design as needed is essential, and the changes
should be justified and recorded for future reference and use. Figure 2.82 provides
some typical parameters used for the Hoek-Brown criterion

ANALYTICAL METHODS

There are three main analytical methods used for mainly coal mine support design:
Beam Theory, Plate Theory and empirical support designs such as that of the US
Corps of Engineers (1980).

• Beam Theory (Timoshenko & Woinowsky-Krieger, 1959)

The stress (σxx) on the roof increases in proportion to the square of the span.
Maximum normal stress is shown in Equation 1:

γ L2
σxx (max) = (1)
2t

The maximum deflection (ηmax) at the centre of a beam is affected even more by an
increase in the span as shown in Equation 2:

γ L4
η (2)
32 Et 2max

Where γ is the unit weight, L is the beam/room span and t is the beam
thickness.
Beam theory assumes only gravity loading, and in reality, the horizontal
stresses have a significant negative effect on the immediate roof stability. The
immediate roof stability is also significantly dependent on the relative orientation
of the rooms and the maximum horizontal stress.
152 Underground soft rock mining

Figure 2.82 Approximate values for the constants (Hoek et al., 1995)

• Plate Theory (Timoshenko & Woinowsky-Krieger, 1959)


Flat Plate Theory can be applied for coal mine intersections. Plate theory considers a
rectangular plate that is clamped at all corners and is based on the following conditions:
• The plate is a straight, flat structural element whose width is at least 4 times
the thickness and whose length is equal to or greater than its width.
Underground soft rock mining 153

• The material is elastic, homogeneous and isotropic.


• Maximum deflection is half the thickness.
• All loads and reactions are normal to the plate.
• When the plate deflects, the central plane remains unstressed.
• Vertical straight lines before flexure remain straight but become inclined to
the vertical.
• The normal stresses in the plane of the plate are proportional to the distance
from the central plane.

For such a plate clamped rigidly at four corners, the maximum deflection occurs
at the centre of the plate and is given by Equation 3:

αqa 4
η (3)
Et 3max

Maximum normal stress is found using the Equation 4:

6β qa 2
σxx (max) = (4)
t2

Where q is the uniformly applied load per unit area, a is the cross-sectional area of
the plate, α and β are coefficients for a uniformly loaded rectangular plate.
A typical intersection has a length-to-width ratio of 1.0, so the coefficients for
α and β are 0.0138 and 0.0513 respectively for this equation when Poisson’s ratio
is equal to 0.3.
Reliance solely on Beam Theory and/or Plate Theory is not recommended due to
uncertainties in the assumptions, as well as in the required input data (Bieniawski,
1989). Geological strata are unlikely to be elastic, homogeneous and isotropic, and
values of Young’s modulus and immediate roof thickness can vary significantly
locally throughout a deposit. The presence of discontinuities and water make it
more difficult to deal with a rock mass using ideal engineering conditions. Rock
Mass Classifications are therefore generally more reliable.
• US Corps of Engineers Support Design
The main part of the empirical method is shown in Figure 2.83 (US Corps of
Engineers, 1980).

COMPUTER MODELLING METHOD

Modelling is the most useful design method because it is fast and parametric stud-
ies can be undertaken, and this is very helpful as input parameters are seldom well
defined and vary anyway. In addition, the model can be calibrated based on in-situ
monitoring results.
Some keys to successful modelling are (Feng et al., 2013):
• The correct model selection based on experience, judgment and testing if applicable
• The determination of representative input parameters
• The setting of the model boundary conditions
154 Underground soft rock mining

Figure 2.83 US Corps of Engineers empirical support guidelines

• Verification of the results by qualitative or quantitative means


• In-situ horizontal and vertical stress magnitudes and principal stress orientations
• Understanding of the problem both in the pre- and post-failure states
• Understanding of the model on a local scale before incorporation to a large global
scale model
• When calibration to a theoretical problem is desired, knowing the assumptions of
these problems and being able to recreate that scenario in the software
• Experience with the program itself and understanding the model’s idiosyncrasies
The goal of any support system is to achieve the equilibrium of the rock mass around
the excavation, retaining as much of the residual rockmass strength as possible. A
concept called the Ground Reaction Curve has been used in the tunnelling industry
for many years (Brown et al., 1983) and can be utilized to roughly estimate how much
support is needed to achieve roof equilibrium. These curves plot the support pressure
against the convergence, as shown in Figure 2.84.
Before excavation, the boundaries have stress equal to the in-situ field stress (Point
A in Figure 2.84). Once the excavation is developed, the reactive/support stress that
is required to prevent additional convergence reduces as the rock structure begins
self-supporting (Point B). A point is reached where serious rock failure and unravel-
ling occurs (Point C), and the self-supporting capacity reduces significantly; the rock
deformation/deterioration accelerates and the required reactive load to create stabil-
ity increases significantly. Ultimately (near Point D), the installed support needs to
carry deadweight (i.e., all the failed ground). Equilibrium for the example shown is at
Point B, where the support curve intersects the ground reaction curve. Using numer-
ical modelling and in-situ measurements of closure in particular, these ground reac-
tion curves and support interactions can be established to find the optimal method of
Underground soft rock mining 155

Figure 2.84 Illustration of a ground reaction curve (Barczak, 2006)

support for a particular roof geology and location (Speers & Spearing, 1996; Barczak,
2006).
The important issues are to:
• Try to understand the problem and the mechanisms that play a role
• Check that the results seem logical (i.e., stress concentrations and the direction of
deformation)
• Try to compare results to reality (difficult for green field applications)
• Update the modelling (and design) as the excavation is being developed

HYBRID METHOD

This involves a combination of methods with back analysis and usually yields the best
solution. Any support design method must, however, be continually reviewed (visually
and with measurements) and amended as necessary and recorded appropriately, list-
ing all observations and assumptions.

2.7.2 Intersection support


The stability of an excavation is generally determined by:

• The rockmass properties


• The stress regime and how it changes due to further mining
• The excavation dimensions (for roof stability the excavation width is critical)
• The presence of water
• Local geological features and properties
• The behaviour of the installed support and when the support is actually installed
(the sooner after excavation, the better)
156 Underground soft rock mining

Intersections are of particular concern regarding roof stability because the effective
span across the diagonal is larger than that of the room, typically 40% more than the
room width. The diagonal span can be effectively even larger because the pillar corners
are typically rounded during their development with a continuous miner and the pillar
corners are also first affected and damaged due to stress concentrations at the pillar
corners. This becomes very significant if the rockmass is relatively weak. In the US,
approximately 71% of all roof falls occur in intersections even though they only account
for 20–25% of the total development. Therefore, a fall is eight to ten times more likely
to occur in an intersection than an entry on a unit length basis (Spearing et al., 2013b).
According to Molinda et al. (1998), numerical modelling and statistical analysis
have shown that although four-way intersections are 1.28 times as likely to fail as
three-way intersections, it takes two three-way intersections to replace a single
four-way intersection. This makes the potential remedy of replacing three-way
with four-way intersections ineffective at reducing the total number of roof falls
in these areas.
The magnitude of the horizontal stresses is also an important parameter, and the
larger the horizontal stress, the less stable the immediate roof, which is particularly
significant in intersections.
In coal mines, secondary support is installed in intersections and usually takes the
form of:
• Cable bolts
• Truss bolts

2.7.3 Types of coal entry support


The most common forms of coal mining support are:

• Resin-anchored rebar (active or passive bolts)


• Cable bolts
• Truss bolts
• Screen/mesh
• Straps

Resin can be introduced using a twin component pump with an in-line mixer or the
most common system that uses a preformed resin cartridge or capsule. The cartridge
is manufactured from a plastic sheet with two separate sections (Figure 2.85): one for
the polyester resin component (mastic) and the other containing the catalyst compo-
nent. The mastic is a mixture of typically a pure fine limestone filler and additives that
accelerate and stabilize the chemical reaction and keep the solids in suspension. The
catalyst component is a mixture of benzoyl peroxide (the actual catalyst), limestone
filler, water, glycol and other additives. After installation in the bolt hole, the rebar is
spun into the hole and ruptures the plastic container, thus mixing the two components
and initiating an exothermic reaction that forms the final resin grout. Figure 2.86
shows an installed fully grouted rebar rockbolt.
There is a misconception that the resin glues the rebar and the rock. It actually forms
a mechanical key against the rock around the hole, thus anchoring the rebar into the
hole efficiently. In a full-column resin-grouted rebar, corrosion resistance is offered
by the resin, but it is not totally effective as the set resin tends to have micro-fractures
Underground soft rock mining 157

Figure 2.85 A resin cartridge construction


Reproduced with kind permission from Jennmar Corp., the owners of DSI in the US

(from the cooling after the exothermal reaction) and subsequent rockmass movement
breaks the resin column, thus permitting water ingress. The spin time and hold time
(if tension is to be applied) are very important. Typically, 40–60 rotations are a good
guide for optimum mixing, and the time to achieve this obviously depends on the
rotational speed of the drilling equipment to be used. The hold time depends on the
gel time of the fastest setting resin component being used.
The annular gap (the difference between the radius of the rockbolt and that of the
bolt hole) is very important with resin bolts to ensure a good performance and should
be ideally kept between 3–6mm if possible. A larger annulus can create mixing issues
and potential shear failure within the grout during loading.
The suppliers of the resin cartridges generally offer a training service for installa-
tion teams that is very helpful and should be regularly undertaken, especially if a new
bolting crew is being used.

2.7.4 Longwall entry (gate) support


Unlike room and pillar entries, longwall gates (entries) are subjected to very high stresses
whose magnitude and direction change significantly as the longwall panel approaches
and passes. The headgate movement is limited, but when the next longwall panel is

Figure 2.86 A full column resin grouted rebar


Source: www.rockbolt.glorysteelwork.com
158 Underground soft rock mining

Figure 2.87 Typical longwall panel showing the entry gates (Einicke, 2009)

mined and the old headgate becomes the tailgate and the deformation is significant, it
can be in the tens of cm. A typical longwall panel and gates are shown on Figure 2.87.
Maingate support consists generally of resin rockbolts, mesh/screen and cable bolts.
Some yielding mechanical or hydraulic props can also be used against the pillar ribs,
especially near intersections.
The tailgate support needs to be able to yield significantly, and additional (supple-
mental) support in the form of free-standing support is installed.
Free-standing supports are set on the mine excavation floor and extend to the roof,
providing a reactive load between the roof and the floor, similar to an adjustable struc-
tural building column.
Free-standing support is frequently used in coal mining for the following main
applications:

• Longwall gate (entry) support especially in tailgates where large deformations are
common
• Where deadweight support is needed
• To hold steel roof beams (steel sets) in place, becoming the support legs

Free-standing support is used as secondary or supplemental support and includes the


following main types:

• Hydraulic props
• Mechanical props
• Pumped cementitious cribs with reinforced plastic or steel outside containment
and reinforcement.
• Cast cementitious cribs
• Timber cribs
Underground soft rock mining 159

Figure 2.88 Some examples of free-standing supports


Photographs taken by A.J.S. Spearing during testing at Pittsburgh Research Lab (NIOSH)

• Timber or steel posts (poles)


• Timber engineered props (elongates)

Examples of some commonly used free-standing supports are shown in Figure 2.88.
According to Spearing (2008), the major requirements for all steel-based adjustable
free-standing supports are:

• Adequate load-bearing capacity and stable deformation (if applicable) for the spe-
cific underground application
• High initial stiffness (early initial strength) and in some applications a pre-load
• The ability for mine workers to safely install the support
• Adequate foot and roof contact area support especially to avoid damaging/punching
the roof (or floor) strata

Other requirements for free-standing supports are:

• Cost effectiveness of the final installed unit, including transportation and instal-
lation costs
• Durability for the application considering the effects of corrosion, etc.
• Low weight
• Ease of transportation
• Ease, speed and reliability of installation
• No ignition sources, such as sparking when installed and carrying load underground
• Minimum interference to men, material and ventilation
• Minimal or no fire potential

Training and ongoing retraining of all the mine personnel involved in the safe and
correct support installation is also critical.

2.7.5 Longwall shield support design


The key performance requirements for longwall shields are:

• Reliability (being maintenance free for hopefully thousands of loading and


unloading cycles)
160 Underground soft rock mining

• Fully automated functions and effective remote-monitored and diagnostic


abilities
• High pre-load and maximum load capacity
• High shield stiffness
• Effective load distribution on the roof and floor
• Small shield tip to coal face distance
• An effective operating height range for site-specific applications
• A short service and spare parts availability and lead time

Seedsman (2002) developed a floor bearing relationship underneath a pillar, and the
same relationship can be used to estimate the bearing capacity of the floor beneath a
longwall shield:

Floor bearing capacity = UCS / 2  4.14 + W / ( 2 / h )

Where:

W is shield width
h is the unit thickness of the immediate floor that has an UCS

Modern longwall shields have unit carrying capacity in excess of 1,500 tonnes.
Moving from a mined-out longwall panel to a new panel stops production, and the
faster the new panel can start mining safely, the better. Most longwall mines therefore
use the same dedicated crew each time because experience can reduce the move time,
and time saved means increased production and reduced costs. The moving of all the
heavy longwall panel equipment is also more hazardous than general production, and
using a properly trained, dedicated and experienced crew can reduce the hazard poten-
tial. The duration can take 2–5 weeks and is dependent on many factors, including:

• The longwall panel length and width


• The longwall seam height
• The equipment available for the move
• The experience of the crew

During the actual move from the mined-out panel to the new longwall panel, the ven-
tilation demand increases and this must be planned for.

2.7.6 Coal bump precautions


Coal bursts involve the violent, sudden and noisy ejection of coal that is over-stressed
and are always very hazardous. Coal bursts require stronger rock as it needs to store
energy to create the burst rather than fail in a quasi-static manner.
The key to any form of seismic support is energy absorption (i.e., work done) so ide-
ally supports must have a high capacity and the ability to yield rapidly without failing
(i.e., maintaining load bearing capacity after the event that is generally associated
with significant closure).
In deep burst-prone coal mines, seismic monitoring using an array of seismometers
or accelerometers is a powerful and critical tool, as is back analysis using computer
Underground soft rock mining 161

modelling after an event to try and determine “critical stress conditions” and then
adjust the mine layout to avoid it in future mining.
Bursts on longwall faces are very difficult to support, and effective strategies are
often limited to pre-conditioning and reducing the mining height.
Multi-seam mining where the inter-burden is relatively small can also cause high
and potentially dangerous stress concentrations.
For potential pillar bursts, wrapping the pillar with screen/mesh is effective as it can
contain the dynamic failure, by safely yielding and absorbing energy. Cable bolts are
effective in sustaining movement and in extreme situations yielding bolts can be used.
The entry roof can be supported using screen/mesh, bolts and cablebolts and steel
straps or steel sets (with yielding legs). Caution must be taken that any steel set legs
are set a distance from the ribs so if the ribs fail violently, the mesh containment can
reduce the chance of shearing (buckling) the legs.
The ultimate solution if the seismic hazard cannot be mitigated or controlled is to
not mine the area.

2.8 Coal mine hazards, safety and risk


2.8.1 Hazard identification and risk management

2.8.1.1 Hazard identification


Any mining operation involves many different activities being conducted simultane-
ously in a natural (so unpredictable) and space-constrained environment. The iden-
tification of hazards is therefore essential in order to protect the workforce, the local
community and the environment. All mining is therefore inherently dangerous because:
• Multiple and often competing activities take place at the same time and in the
same working area.
• Employees continually interact with moving parts and equipment.
• Equipment and processes are expensive, so breakdowns are serious and costly.
• Unplanned events occur routinely, and they can be significant.
• The probability of loss is high.
• Costs of an accident or loss are high and often serious.
The hazards most associated with coal mining are:
• Waste tailings dam stability
• Methane and coal dust explosions
• Spontaneous combustion
• Fires
• Floods
• Coal dust inhalation
• Falls of ground (roof and rib)
• Coal seismicity
• Unplanned pillar failure
• From mobile equipment
• Electrical problems (only permitted equipment can be used, and it must be kept
in compliance at all times)
• Tripping and falling
162 Underground soft rock mining

Hazard identification must be based on knowledge and compliance with relevant reg-
ulations, management commitment and sound, regular and effective training. The
workforce must be empowered so they have the confidence to raise all safety-related
issues to management immediately and have these taken positively and resolved.
Hazard identification needs to be conducted before mining operations or spe-
cific new activities are undertaken as this process needs to be always proactive. Any
changes to existing operations, tasks or procedures must be preceded by a full hazard
identification and procedure review.
Tools or actions that can be used for hazard identification include:

• Records of accident and incidents from the mine or other similar mine sites
• Mine routine checklists
• Involving relevant mine operators to identify potential problems and hazards
• Federal, state or local relevant databases

2.8.1.2 Risk management in coal mines


Risk management is covered in more detail in Section 4.6. Generally, risk is related to
the expected losses which can be caused by a risky event and to the probability of this
event happening:

• The more severe the loss and the more frequent the event, the worse the overall risk.
• Measuring risk is often difficult; rare failures can be hard to estimate, and any loss
of human life should be considered irreplaceable.

The engineering definition of risk is:

Risk = ( probability of an accident ) × ( losses per accident )

Risk must be effectively eliminated or mitigated to:

• Prevent incidents or accidents


• Ensure similar recurrences are avoided
• Reduce medical costs
• Reduce workman’s compensation premiums or the local equivalent
• Reduce insurance premiums
• Keep employee morale up and keep the personnel safe
• Maintain a more positive image with the local community, other stakeholders and
the media

The following steps are taken to eliminate, control or reduce risk:

• An assessment must be conducted that includes reviewing accidents, injuries,


first-aid cases and loss records.
• Action must be taken to manage the risk (and recorded).
• Follow-up observations and audits must be conducted.
• Report findings and, if necessary, the corrective actions that were taken.

Figure 2.89 shows a typical hierarchy of control for a risk assessment program.
Underground soft rock mining 163

Figure 2.89 The risk assessment hierarchy of controls


National Institute of Occupational Safety and Health, 2015

It is important that the risk management system used is all inclusive as safety must
be the first and most important consideration in mining. It needs to be regularly revis-
ited and updated based on continuous and routine reviews of all incidents and acci-
dents. Avoidance of a recurrence is most critical, as a risk-free environment is not
really possible.
Risk management involves the following steps:

• The entire mine staff and owners must first be committed to it.
• An inclusive Risk Management Committee and Sub-Committees must be formed,
giving participants authority, relevant training, recognition and time to under-
take the necessary work.
• The risks (likelihood and consequence) must be identified and correctly cate-
gorised, hopefully starting with the most serious potential risks on the specific
mine.
• The risks must be evaluated.
• Methods, procedures and rules to eliminate or mitigate the risks must be devel-
oped and continually reviewed and updated as needed.
• These must be communicated to all affected staff and local community
regularly.

Risk management should be consistent across the globe, but it tends not to be, as
the social and living standards in a region or country also affect what is considered
acceptable risk and what is not. Mining companies with international operations,
however, try to be consistent across all their mines, which is good.
164 Underground soft rock mining

2.8.2 Coal mine safety training


Coal mine training is very important, especially safety-related basics such as don-
ning a self-rescuer or not taking flammable contraband underground. The training,
however, must be made interesting and the workforce must understand the need and
relevance of it. It is simple to correctly don a test self-rescue unit in a training room,
it is a completely different situation to don it during an emergency where time is
critical and seconds can make the difference between life and death.
The introduction of virtual reality (VR) to mine training has the potential to make
training real in the same way as flight simulators are effectively used to take pilots
through challenging flight situations safely but realistically. VR is, or can be used, for
the following training applications:

• Electrical fault finding and safety-related maintenance procedures


• Rock-related hazard identification
• Mine rescue training
• Safe barring practices
• Much of the new miner training for new employees to mining
• Equipment operation and driver training
• General safety training (e.g., donning procedure for a self-rescuer)

Audits are an effective way of identifying areas that need improvement or finding root
causes of hazards.
Audits should be conducted at various levels. Ultimately, industry or state-wide
audits have the following objectives, according to Lynch (2011):

• Assess whether the system for managing health and safety at mining oper-
ations includes all matters, plans and procedures required under relevant
legislation.
• Provide feedback to industry on the extent to which the systems for man-
aging health and safety at mining operations comply with the relevant
legislation.
• Identify industry-wide problems and issues in achieving compliance with the
relevant legislative provisions and suggest strategies to address them.
• Establish baseline data on industry compliance with the legislative provisions to
enable trend analysis in future compliance audit programs.

2.8.3 Coal mine hazard and safety monitoring


A Bow Tie analysis is one of the simplest and effective methods of assessing risk and
handling it and can be undertaken with little specific training. It combines an Event
with the potential Causes and Consequences in an easy to visualize way as shown
in Figure 2.90. These can be done manually or using appropriate software packages
which can be very useful.
Figure 2.91 gives a Bow Tie Analysis example for a methane explosion in a coal
mine.
Underground soft rock mining 165

Figure 2.90 The skeleton of a Bow Tie Analysis

Figure 2.91 Bow Tie Analysis for a methane explosion in an underground coal mine

2.9 Coal mine rehabilitation and reclamation


Acid Mine Drainage is probably the single most critical environmental concern in
coal mine rehabilitation and reclamation. Mines therefore need to focus on:

• Environmental regulation compliance


• Relevant permits (such as water discharge and burning)
• The involvement of all the various relevant government agencies [such as the
US Army Corps of Engineers, US Fish and Wildlife and the Environmental
Protection Agency (EPA) to mention some in the US]
• Highwall, slope and subsidence remediation
• Gas venting and well remediation
• Endangered species protection
• Funding of the reclamation
• Meeting all bond requirements so they are not lost due to non- or only partial
compliance
• The current and future spontaneous fire potential (both on surface and underground)
166 Underground soft rock mining

Mapping (using photogrammetry, LiDAR, Bathymetry, etc.) is a useful tool to aid the
process.
Liners (or similar such as blankets, geofabrics and geotubes) are now commonly
used for most ash ponds, and if correctly built, they limit the leaching caused by
percolation resulting from rainfall. Water overtopping is a serious potential issue,
and the design should be such that this can be avoided even under extreme weather
events.
The goals of any mine-related closure must be:

• To return the environment to an acceptable long-term condition based mainly on


regulations but also taking account of the affected local community and private
owner (if applicable)
• To repurpose or reuse as much of the waste products as possible

Effective and transparent involvement with the local community throughout the life
and closure of the mine is essential.
Mine closure must be done carefully and as socially responsible as possible because
typically the closure is associated with significant employee reductions, and ultimately
total reduction, and this has a negative effect on the local economy. The land must
also be returned to its original condition, or that mutually agreed to by all stakehold-
ers. Local regulations are the minimum that a mine under closure must meet, but
more should be expected considering the negative local economic impact (employees
losing their jobs, the local community losing a large economic input from mining
employees, etc.) and the local community expectations. Legislation currently tends to
only address direct environmental impacts, which whilst very important is far from
the only consideration.
A Mining Minerals and Sustainable Development report by the International
Institute for Environment and Development (2002) recommended that mines under-
take this additional work associated with the social aspects of mine closure:

• Create effective management systems (or engage skilled consultants and con-
tractors) to continually review end-of-mine life plans, to take necessary actions
to strengthen them and to monitor them throughout the mine life and closure
process
• Focusing on whether existing plans fully address the end-of-life environmental,
social and economic conditions for affected communities and their potential
enhancement, care and opportunities for displaced workers and the implications
for government and other stakeholders at all levels

2.9.1 Mine access closure


Mine access and plant closure must be safely undertaken based on sound engineering,
economic, environmental and local community considerations. Safety of the workers
involved in the closure processes are of paramount consideration. Knowing where all
workers are located throughout the working shifts is essential, and in general, workers
should not work alone and everyone should be able to communicate with each other
using mobile phones or radios.
Underground soft rock mining 167

Shafts, ramps and declines must be decommissioned and as much as possible of the
above ground structures removed from the site, if they are not to be utilized by the
local community. This is so that:

• The hazards caused by access to humans and animals are reduced.


• Potential hazardous gas emissions are stopped.
• Water contamination is reduced.
• The land is fully rehabilitated to its original contours where possible, sold
without further liability or developed as agreed to by all parties (especially the
regulator).

According to the US National Resources Conservation Service (2005) – Mine Shaft


and Adit Closure, the effective methods to close such underground access are by:

• Filling
• Plugging
• Capping
• Installing barriers, gating and/or fencing

Generally, areas that are fenced off need to be regularly inspected and maintained to
ensure no unauthorized entry after closure.
In the report, it was further recommended that:

Shafts and adits shall be cleaned of all trash, debris, metal, timber, wire and
other materials that could hinder an effective designed filling or sealing. The fin-
ished surface of the filled or plugged shaft or adit shall be graded to provide free
drainage away from the opening and vegetation established in accordance with
standards. All materials removed shall be disposed of by burning or burying at
approved sites or transported to approved landfills.

This is mainly for shallower mines, and in deeper shafts, suspended platforms can be
used to help the access sealing.

2.9.2 Coal mine waste rehabilitation


Most coal mines especially in the US use conventional surface impoundments to dis-
pose of washing plant waste. Rehabilitated land uses include:

• Forests
• Grazing pastures
• Crop fields
• Nature parks

Generally, the legal requirement is to return the land and infrastructure to that prior
to mining. A better approach may be to consider the rehabilitated land use, based on
input from the private owner or the local community.
168 Underground soft rock mining

2.9.2.1 Avoiding coal mine waste


The most environmentally friendly solution is to not create surface coal waste and
rather use it all to backfill the coal mine in the vicinity. Backfilling using coal waste
is not a novel means for waste disposal throughout the world, and has been used fre-
quently in China, Germany and Poland. Unless a thermal power plant is located close
to the coal mine, there is not enough washing plant waste (if such a plant is used) to
fully backfill all coal mine workings. If complete backfill is essential for environmen-
tal reasons, then another source of backfilling material will be needed to augment the
coal waste, such as sand, other industrial waste or quarried waste.
Many mines, however, prefer to let the subsidence and its associated damage occur
then rehabilitate appropriately, especially if there is farmland above and little ground
water or infrastructure.
One of the overlying and remaining issues with underground disposal is the poten-
tial contamination of the water table in the long term, mainly from water runoff
through the placed backfill (Earthjustice, 2009). This is only a potential hazard if
the coal mining depth is near that of the deepest groundwater. To mitigate this, the
backfill needs to be placed as a high density paste or dry. If water runoff does occur
where it can potentially contaminate the groundwater, then it should be collected and
treated underground before it is released, if released.
Backfilling can also provide benefits to active underground coal mines depending on
the mining method, such as stress reduction of the longwall face in a deep coal mine.

ROOM AND PILLAR COAL MINING

The main advantages that could be derived from backfilling room and pillar panels
after they have been mined are:

• Potentially improved extraction as the stand-up time of mined panels will be


short, so pillar dimensions could be reduced, as once a panel is mined, it will be
backfilled.
• Better ventilation control (including reduced methane ingress) during and after
mining by effectively sealing mined panels.
• Effective blasting seals to reduce the chance of methane problems in old mined-
out areas.
• No long-term trough or pit subsidence from collapses along rooms or at intersec-
tions respectively caused by weathering (and even flooding of old mines).

LONGWALL MINING

Longwall coal mining with high-density backfill has been common as mentioned
for several decades in Germany and Poland. Its main purposes were to reduce sur-
face subsidence, protect surface and groundwater and help reduce stresses. This has
proved effective, and China is now routinely using mainly solid (dry) backfill mining
(Zhang et al., 2019) under sensitive infrastructure such as water, highways, railway
lines and buildings.
Appropriately designed backfilling is a proven technology for mitigating water
damage and surface subsidence. Depending on the relative location of the mine and
coal waste impoundments, additional waste or other material (such as sand) may be
required for effective high percentage backfilling.
Underground soft rock mining 169

Figure 2.92 The plant decommissioning and remediation process (EPA, 2016)

2.9.2.2 Closure and remediation of coal mine waste (EPA, 2016)


The closure and remediation of coal washing plant, ash ponds or dumps (impound-
ments) vary considerably based on the applicable local regulations and the size, loca-
tion, topography, climate and infrastructure where it is located. Typically though,
the process involves: dewatering, detoxification then cover with soil and finally
vegetated.
The procedures to be followed are shown in Figure 2.92 (EPA, 2016), and Figures 2.93
and 2.94 show remediation work on a coal tailings dump.
Considerations for a pond closure typically include the following:

• Safety during and after the closure – Safety must always be paramount, and a
thorough site-specific risk assessment must be undertaken before the start of the
planning process. All parties must be appropriately trained and regular reviews
and meetings held. The site must be secured to keep the public away, and the pro-
cedures and end objective should have been discussed with the local community
and private land owner, if applicable (and agreed to by all).
• Regulation compliance – At all times during the remediation, relevant regulations
must be met (and preferably exceeded). Water discharge monitoring and ensuring
it complies (and exceeds) the minimum quality levels required. The effects of sig-
nificant rain events must be planned for, and adequate handling facilities (such
as sumps, etc.) must be provided to handle runoff and ensure it is treated before
release from the mine site.
• Total costs – Depending on the relevant regulations, the cost of remediation may
be covered via a bond that was provided prior to mining (usually as part of the
permit application), but this is not always the situation. Costs can be significant.
• Pond dewatering – This is a critical and often a costly part of the remediation
process. Excess water must be removed such that the ash will not liquefy and can
be safely stacked.
• Ash particle size distribution – Particle size distribution and mineralogy are the main
factors that determine the shear strength of the ash, which is critical to the stable long-
term stacking of the dewatered ash.
• Dust control – Effective dust protection must be made available for all site workers
and actions taken to limit dust affecting the local community. Coal ash is generally
(but not always) acidic so prolonged contact with unprotected skin should be avoided.
170 Underground soft rock mining

Figure 2.93 Coal waste levelling at the start of reclamation


Photographs reproduced with kind permission from John T. Boyd Company
Underground soft rock mining 171

Figure 2.94 Coal waste reclamation during levelling and before soil cover
Photographs reproduced with kind permission from John T. Boyd Company

• Surface water impacts – Coal mine water often contains high levels of iron, man-
ganese and sulphates in addition to lower levels of other contaminants such as
mercury and cadmium. Waste coal water treatment can be chemically or passively
based. Passive water treatment involves less chemical usage, can be cheaper but
typically requires more land. The larger land space is needed as the wastewater is
ponded and kept in place until the level of contaminants has fallen to acceptable
levels. Passive systems typically include one of the following types (or a hybrid):
• Aerobic wetlands – Artificial aerobic wetlands have a large surface area and
are filled with cattails (a hardy and fast-growing plant with very long and
narrow leaves) or other hardy plant species (such as phragmites). They treat
alkaline waste water, which is not the most common. The metals precipitate
out as oxides or hydroxides.
• Organic substrate wetlands – These use anaerobic processes to precipitate
metals.
• Anoxic limestone drains – Anoxic means that the general environment is kept
basically oxygen-free. This consists of a bed of limestone that increases the
alkalinity of the water.
Metal removal usually involves hydrolysis and oxidation.
• Site-specific constraints – Each site is unique and can present unique challenges.
Considerations include: the topography as this can influence the stability and
path of drainage water for sump and treatment location, the climate especially
rainfall, proximity of surface water and any local communities.
• Long-term monitoring and maintenance – Water egress from the remediated ash
pond must be collected and monitored for pH, total dissolved solids (TDS) and
total suspended solids (TSS). TSS content is generally the parameter that is
172 Underground soft rock mining

Figure 2.95 Typical closure stages for a coal ash pond (Johnson & Nilsson, 2014)

regulated. With new technology, this monitoring can now be done in real time.
Usually, TSS control involves sieving to remove the larger material and debris
following the addition of a coagulant.

Mountain top mining significantly complicates the processes and costs needed for
mine reclamation.
A simplified outline of an ash pond remediation is shown in Figure 2.95.

2.10 Green mining developments

2.10.1 Water conservation and subsidence control


Subsidence and hydrology impacts occur due to underground mining operation to
various extents, which often bring about adverse changes to surface topography,
ground water and surface water.
The adverse effects on water conservation and subsidence on other man-made sur-
face structures and other features are relatively well known and studied, but the envi-
ronmental impacts related to subsidence and hydrology at underground mines are
typically not as well understood or considered.
Groundwater is affected by mining and drainage gradients that may be altered by
mining, which causes fracturing and weakening in general. Rock may become weak-
ened by weathering.
Where surface water is present, it may migrate easily into mining-induced fractures
and fissures in the strata and even into the mine area and may induce subsidence.
Another mining-related activity that can also create subsidence is the dewatering
that may be needed to facilitate safe underground mining. Water removal can also
cause the formation of cavities (which were once filled with water), and these may
result in subsidence as the hydrogeological properties of the associated strata are
adversely changed.
Underground soft rock mining 173

Subsidence caused by mining is not a new problem. In 1556, Georgius Agricola


noted in De Re Metallica, his classic work on mining and metallurgy, that the Italians
had forbidden any mining under or near the extensive vineyards and fields of the prime
agricultural regions. This was because of the negative impacts of mining subsidence
and the degraded water quality caused by mining.
Subsidence is both a natural and a man-made phenomenon associated with a vari-
ety of processes such as compaction of the natural soils, ground water level changes,
melting of permafrost, liquefaction and crustal deformation, withdrawal of petroleum
and geothermal fluids and general mining (Soliman, 1998).
Singh and Dhar (1997) report that, “Subsidence is an inevitable consequence of
underground mining – it may be small and localized or extend over large areas, it may
be immediate or delayed for many years”.
Fejes (1986) call subsidence “a natural result of underground mining”, and goes
on to state that, “When a void is created nature will eventually seek the most stable
geologic configuration, which is a collapse of the void and consolidation of the over-
burden material”.

2.10.1.1 The effects of subsidence and the causal factors


Singh and Dhar (1997) define mine subsidence as “ground movements that occur due
to the collapse of overlying strata into mine voids”, which expresses itself in three
major ways:

• Cracks, fissures or step fractures


• Pits or sinkholes
• Troughs or sags

A number of geologic and mining parameters can affect the magnitude and extent of
subsidence, including:

• Thickness of extracted materials


• Overlying mining areas
• Depth of mining
• Dip of mining zone
• Competence and nature of mined and surrounding strata
• Near-surface geology
• Geologic discontinuities

Additional factors effecting subsidence have also been identified (Singh and Dhar,
1997):

• Fractures and lineaments


• In-situ stresses and stress changes as mining continues
• Degree of extraction
• Surface topography
• Ground water
• Mine area
• Method of mining
174 Underground soft rock mining

• Rate of mining advance


• The use of backfilling
• Time
• Structural characteristics

2.10.1.2 Subsidence prediction


Underground mining can result in surface subsidence, and the main issues are:

• The magnitude of the vertical subsidence


• The lateral extent of surface subsidence
• The time over which the subsidence will occur
• The nature of the subsidence (i.e., continuous or sporadic/episodal)

Five principal methods of predicting mining subsidence have been developed (Whittaker
& Reddish, 1989):

• Empirical relationships – These are a broad category that includes numerous meth-
ods. Empirical methods are based on field observation and experience and are gen-
erally applied to regions where adequate empirical data are available. Empirical
methods for prediction of subsidence consist of graphical, profile function and
influence functions that are constrained by the availability of observed data.
Empirical methods are quick, simple to use and relatively accurate but should
be used within the data parameters used to establish the specific empirical rela-
tionship. These methods provide satisfactory results and are widely used in coal
mining.
• Profile functions – The Profile Function Method is a curve-fitting method using the
measured subsidence profiles in a particular mining area. There are many profile
functions developed empirically for nearly all major coalfields in the world. These
include for the United Kingdom, the National Coal Board’s (1975) Subsidence
Engineer’s Handbook, and for the United States, it is outlined in Karmis et al. (1984)
and Agioutantis et al. (2014). These are generally only applicable to single panels,
since they assume the profiles to be symmetrical and fail to recognize the way in
which subsidence shapes are modified over adjacent and previously mined areas.
• Influence functions – The Influence Function Method was first proposed in 1931
and has been further developed and refined since then (Zenc, 1969). This method
is now widely used in the world and can be applied to a wide range of mining
geometries, but it is more difficult to calibrate than the Profile Function Method.
• Analytical/numerical models – Numerical methods, particularly Finite Element
Analyses, have been found useful for better understanding and analyzing
complex geology and mining. Computer-based solutions are useful because
they can model rockmass properties reliably. This is a numerical method and
it can simulate inhomogeneity, bedding planes, joints, faults anisotropy and
relevant boundary conditions. Input sensitivity analyses can be readily con-
ducted, but the modelling and analysis of results need skilled and experienced
users.
• Physical models – Scaled physical models have been used to model subsidence and
also to help calibrate numerical models.
Underground soft rock mining 175

• Phenomenological methods – Phenomenological methods are based on modelling


principles, which use mathematical representation of idealized materials with the
application of continuum mechanics.

According to Singh (1986, SME), “the (phenomenological) procedures had not


achieved much success to date, mainly due to the difficulty of representing complex
geologic properties of the strata in simple mathematical terms”.

2.10.1.3 Coal mining subsidence


Coal mining subsidence is highly dependent on the mining method.
• Room and pillar-induced subsidence
Room and pillar coal mining is typically suitable and economic in shallow coal depos-
its (<300m) with strong immediate roof rock and variations in coal seam continuity/
quality. Room and pillar mining is effective as the overburden load is carried by the
pillars and designed using the Tributary Area Theory. The roof is artificially sup-
ported using installed rockbolts often with some free-standing support for the useful
duration of the room and pillar panel. Sometimes pillars are partially recovered by
secondary pillar extraction in a retreat fashion, which increases the loads on the pillar
system. Since extraction is relatively close to the surface, it is easy for subsidence to
reach the surface, typically several years after extraction, caused mainly by weather-
ing (water). These effects depend on the lithological setting, related geomechanical
features and the site specific mine design.
Sinkhole subsidence is common in areas overlying shallow room and pillar mines.
Sinkholes occur from the collapse of the mine roof typically at an intersection (which
has a wider effective span) into a mine opening, resulting in caving of the overlying
strata and a depression or “chimney” type hole in the ground surface. The majority
of sinkholes usually develop where the amount of cover (vertical distance between the
coal seam and the surface) is less than about 50m. This type of subsidence is generally
localized in extent, affecting a relatively small area on the overlying surface. However,
structures and surface features affected by sinkhole subsidence tend to experience
extensive and costly damages, sometimes in a dramatic fashion. Sinkhole subsidence
has been responsible for extensive damage to numerous homes and property through-
out the years in the US.
Aside from pillars and rooms being wrongly mined, the main potential for localized
subsidence is time-dependent, as a result of weathering. Pillars, the floor and roof can
weaken over time especially as mines tend to flood after production ends.
• Longwall-induced subsidence
Longwall coal mining subsidence and longwall mining with backfill for subsidence
control are covered in Sections 2.5 and 2.6.

2.10.1.4 Potential hydrologic impacts of underground mining


Underground mine openings can intercept and convey surface water and ground-
water via mining-induced fractures. When mining below the water table, mining
176 Underground soft rock mining

voids create paths that can induce groundwater to migrate into the openings from
water-bearing rock above. The result is the dewatering of these water-bearing rock
units via drainage through fractures that join the mined volumes with the water-bear-
ing strata. There is also the potential for impacts to more remote water-bearing units
and surface water bodies depending on the degree of hydrologic communication.
The extent and severity of the impact on the local surface water and groundwater
systems depends on the:

• Extent of the mining


• Depth of the mine
• Lithology, topography, geology and hydrology

In addition, the amount and extent of mine subsidence governs the impact of under-
ground coal mining on surface and groundwater.
In the flat-lying sedimentary rocks of southwestern Pennsylvania, underground
mining is routinely accompanied by rock fracturing, dilation of joints and separa-
tion along bedding planes. Rock movements occur vertically above the mine work-
ings and at an angle projected away from the mined-out area. Mining-induced
fracturing within this angle can result in hydrologic impacts beyond the margins
of the mine workings. The zone along the perimeter of the mine that experiences
hydrologic impacts is said to lie within the “angle of dewatering” or “angle of
influence” of the mine. Angle of influence values of 27–42° have been reported for
the coalfields of northern West Virginia and southwestern Pennsylvania (Carver
& Rauch, 1994).
These changes to the rock mass can change the water-transmitting capabilities of
the rock by creating new fractures and enlarging existing fractures. This typically
results, at least temporarily, in detectable changes in permeability, storage capac-
ity, groundwater flow direction, groundwater chemistry, surface-water/groundwater
interactions and groundwater levels. Depending on the ratio of overburden-to-seam
thickness and the type of mining, measurable surface subsidence may occur. As pre-
viously discussed, this surface movement ranges in type from broad troughs approxi-
mating the area of coal extraction (typical of longwall mining) to complete collapse of
overburden from the mine to the surface, e.g., sinkhole subsidence (generally associ-
ated with shallow room and pillar mining).
The two common underground coal mining methods have distinctly different
impacts on local water resources. The impacts of room and pillar subsidence tend to
be localized, random and often take a long time to manifest, whereas longwall subsid-
ence tends to be immediate, pervasive, systematic and ultimately predictable (Booth
et al., 1998).
According to Parizek and Ramani (1996), the enhancement of the overburden
hydraulic conductivity due to mining is neither uniform nor well-defined. Predicting
impacts is difficult and there is no such thing as a “typical” hydrogeologic setting or
mine site.
• Potential Impacts on Streams and Surface Waters
The impacts of underground mining on surface waters can range from no noticea-
ble impact to appreciable diminution, ponding and/or diversion. The formation of
Underground soft rock mining 177

subsidence-induced cracks, surface depressions and/or sinkholes at the bottom of, or


adjacent to, surface water bodies such as streams, ponds and lakes can lead to com-
plete or partial loss of water due to leakage to the underlying strata. The resultant
changes in surface slope can adversely impact drainage along irrigated fields, canals,
sewers and natural streams (Bhattacharya & Singh, 1985).
Room and pillar mining is generally less disruptive to nearby surface waters than
high-extraction methods. Individual openings have only minimal localized draining
impacts due to self-supporting roof members which span the opening to form a com-
pression arch, with the support pillars serving as abutments. This “pressure arch” lim-
its not only the deformational, but also the hydraulic influence of the opening (Booth
et al., 1986). As additional entries are driven, the resultant network of intersecting
drains act as a planar underdrain inducing downward leakage from overlying units.
However, due to its built-in system of support pillars and limited mining-induced frac-
turing, significant drainage is typically limited to near-mine units.
Many detrimental subsidence-related issues with room and pillar mining take years
to occur as weak coal pillars, roof and floor deteriorate over time, often due to water
ingress after mining. Deteriorating or under-sized pillars that fail over time result in
vertical extension of mine-induced fracturing. Dewatering impacts under these con-
ditions can reach to a few hundred feet above the mine collapse areas. Rauch (1985)
provides the following description of the dewatering impacts of room and pillar min-
ing in the north central Appalachians, in the US:

Typically the greatest groundwater inflow rates occur near the working face of
the mine where groundwater is being drained from storage, especially from frac-
tures in mine roof rocks. In older mine sections, long term groundwater recharge
to the mine is under more or less steady state conditions, originating ultimately
from infiltration of precipitation or surface water. This water typically enters the
mine along rock fractures that intersect the mine ceiling, especially along vertical
fracture zones. Groundwater inflow is especially great in areas of mine ceiling
collapse due to the leaving of too little roof rock support or to weak ceiling rock
where fracture zones intersect the mine.
This drainage to room-and-pillar mines dewaters some overlying aquifers. The
extent of this drainage is best determined from studies of water wells and springs
overlying the mines. In general, significant dewatering extends to 6 to 30 m (20 to
100 feet) vertically above drained room-and-pillar mines, but is usually restricted
to within about 12 m (40 feet) vertically of these mines.
“Localized, significant hydraulic impacts of deep headings and un-collapsed
room-and-pillar mines will be seen in shallow aquifers only in areas (such as frac-
ture zones) where vertical hydraulic connections are naturally high or where the
mine itself is very shallow” (Booth et al., 1986). Shallow room-and-pillar min-
ing [within 61 m (200 feet) and particularly within 30 m (100 feet) of the surface]
drastically increases the likelihood of significant impacts to surface waters. This
results from the mining’s proximity to shallow, open fractures and unconsoli-
dated surface material.

Hobba’s (1993) report on room and pillar mining in northern West Virginia, US
found that, mining and subsidence cracks increase hydraulic conductivity and
178 Underground soft rock mining

interconnection of water-bearing rock units, which in turn cause increased infiltra-


tion of precipitation and surface water, decreased evapotranspiration, and higher
base flows in some small stream. Both gaining and losing streams were found in mined
areas.
In deep settings, the impacts tend to be minimal. Bruhn and Speck (1986) reported
the following impacts from room and pillar mining conducted beneath 183m (600 feet)
of overburden in northern West Virginia, US:

Structurally, the overburden strata were little affected by the introduction of


entries and cross cuts into the panel. Subsidence was virtually nil. Piezometric lev-
els remained at their pre-mining elevations except (presumably) near mine level.
Measured piezometric level variations were minor and no more than might be
expected from seasonal variations.

• Potential Impacts on Wells and Springs


Hobba’s (1993) report showed that wells and springs in proximity to room and pillar
mining have the potential of being adversely impacted. Commonly, the mechanism
is direct draining of groundwater to the mine. Generally, where the support pillars
are stable, these impacts are localized. Dewatering typically extends to 6–30.5m
(20–100 feet) above the mine workings. Wells that terminate at depths greater than
31m (100 feet) above the mine roof are generally safe. In cases where support pillars
fail, additional subsidence may result in more extensive fracturing. In these instances
impacts may be up to 61m (200 feet) or even 92m (300 feet) above the mine roof.
Subsidence impacts may be extended where mining is close to vertical fracture zones.
• Other Causes of Impacts on Structures and Water Supplies
There are many factors that result in damage to structures other than mine subsid-
ence. Most structures have some degree of damage even prior to mining. Many prop-
erty owners are not fully aware of the condition of their home and property since they
do not generally conduct thorough structural inspections. It is therefore important
for the mine operator to assess this prior to any mining as it can save money and help
separate genuine claims from perceived or even fabricated claims.
The public sometimes like to take advantage of nearby mining and attribute all
damage to the mining subsidence though the damage may be partially of totally
unrelated.
Some of the more common reasons for structural damages other than mine subsid-
ence are:

• Settlement due to drying of soils or the weight of surface loads


• Landslides and soil creep
• Shrinking and/or swelling of soils
• Freezing and thawing of soils
• Surface and subsurface erosion
• Change in the water table level
• Poor building construction methods, products or design
• Structural deterioration due to age, lack of maintenance or misuse
• Structural movements
Underground soft rock mining 179

2.10.1.5 Impacts of mining-induced subsidence on surface


The following are general observations that could indicate damage caused by mine-
induced subsidence (especially longwalling):

• Ground cracks generally associated with the tensile zone of the longwall subsid-
ence profile
• Ground cracks parallel to the advancing longwall face caused by the sudden sub-
sidence due to longwall mining, but these cracks tend to close as the face contin-
ues to advance
• In areas that are prone to landslides, it is common for slips to occur, particularly
in areas within the tension zone
• Changes in surface contours may cause low-gradient streams to pond and
f lood in the longwall subsidence trough area, sometimes creating seasonal
wetlands

Damages to structures are generally classified as cosmetic, functional or structural.


Cosmetic damage refers to slight problems where only the physical appearance of the
structure is affected, such as local cracking in plaster or drywall. Functional dam-
age refers to situations where the structure’s use has been adversely affected, such as
jammed doors or windows. More significant damages that impact structural integrity
are classified as structural damage. This would include situations where entire founda-
tions require major repairs or replacement due to severe cracking of supporting walls
and footings.

2.10.2 Repurposing coal and thermal waste


Coal mining has the advantage that all the waste produced from mining can be back-
filled underground. This is not the case with metal mining, where the very low recov-
ery and swelling caused by rock breaking create much more waste than can be used for
backfilling. The main waste issues from coal are caused by thermal power generation
where fly ash (a useful product) and bottom ash is created. The thermal coal waste is
also acid generating because of the pyrite and often contains traces of mercury, chro-
mium, lead and arsenic.
According to the US Environmental Protection Agency (www.epa.gov/coalash/
coal-ash-basics), typical combustion products are:

• Fly Ash, a very fine, powdery material composed mostly of silica made from the
burning of finely ground coal in a boiler.
• Bottom Ash, a coarse, angular ash particle that is too large to be carried up into
the smoke stacks so it forms in the bottom of the coal furnace.
• Boiler Slag, molten bottom ash from slag tap and cyclone-type furnaces that turns
into pellets that have a smooth glassy appearance after it is cooled with water.
• Flue Gas Desulfurization Material, a material left over from the process of reduc-
ing sulphur dioxide emissions from a coal-fired boiler that can be a wet sludge
consisting of calcium sulphite or calcium sulphate or a dry powered material that
is a mixture of sulphites and sulphates.

It is very important that these trace toxic contaminants are kept out of the water system.
180 Underground soft rock mining

Typically about half the waste (the fine precipitated fly ash) has value as a Portland
cement extender or to make gypsum board. Transportation from the source to the end
user can create a cost constraint.
Currently the main uses for bottom ash, which are even more transportation
dependent, are:

• Backfilling
• Partial replacement for aggregates (Thi et al., 2019)
• Brick making (Naganathan et al., 2012)
• Ceramics, water treatment, fire-proof products and soil enhancement (Jayaranjan
et al., 2014)

Backfilling would seem to hold the most potential to dispose of large quantities of the
bottom ash safely and beneficially.

2.10.3 Coal rockburst precursors, mitigation and control


Coal bursts (rockbursts) are a sudden and dynamic rock failure associated with the
sudden release of stored strain energy. This represents a serious risk for safety and
productivity particularly as there is often no real warning of such a specific hazardous
event.

2.10.3.1 Factors that can cause coal bursts


The following factors contribute to the rockburst hazard potential in coal mines
(Mark & Gauna, 2015):

• Mining depth – Generally there is low potential for coal bursts in seams less than
300m deep and the potential increases as the seam depth increases.
• Pillar design – The problem is generally associated with pillars that are too small
to safely distribute the applied load but too big to be a yielding pillar (an interme-
diate zone).
• Multiple coal seam interaction (if applicable) – High stress interactions can occur
between seams where the interburden is limited. Remnants on the one seam can
create additional problems on the other seam.
• Geology – Strong thick rock layers in the immediate floor or the roof can also act
as stress concentrators that can induce sudden and rapid failure.
• Mining layout and practice – The mining method and layout can be a significant
source of coal bursts, especially with room and pillar mining with secondary
pillar extraction. Retreat longwall mining, which is the most efficient longwall
sequence, has the disadvantage that it produces high face and pillar stresses
around the working area where people are present. The width and mining height
of the longwall can also be a contributing factor, as can the shearer cutting
width.
• A prior history of rockbursts and high stress situations – These can indicate an
obvious and increasing risk so mitigation methods must be implemented and their
success monitored.
Underground soft rock mining 181

Gas outbursts can also be a potentially serious problem that can cause similar coal
bursts.

2.10.3.2 Coal burst precursors


Seismic systems using geophones or accelerometers are useful to detect potential
problem “hot spots” and, after an event, can show rescuers where they need to focus
their rescue efforts initially.
Researchers (Kim et al., 2015; Sun et al., 2021) have been using acoustic emission
(AE) and infrared radiation (IR), which are being studied as a potential precursor to
violent failure. The energy released by stress-induced rock damage measured using
the AE and IR anomalous phenomena are caused by the development of internal
cracks or defects of rock induced by increasing stresses. The AE and electromag-
netic radiation produced in the rock includes effects in the infrared wave band and
are as a result of physical effects (as sound and heat) and geochemical effects (Zhao
& Jiang, 2010).

2.10.3.3 Coal burst mitigation and control


Rockburst supports need to be able to absorb energy and the advancing face and pil-
lars are the most at risk. The correct pillar sizes are essential as they should be either
large enough to safely carry the stresses generated or small enough to safely and stably
yield in a controlled manner. The strength of the roof and floor must also be consid-
ered in any pillar design.
The mining layout can often be optimized to reduce the maximum stresses and
keep adverse geological features (such as faults and stiff dykes) from failing. Computer
modelling of different layouts can be qualitatively used and are very effective. The
selection of representative input parameters is critical, and some form of model vali-
dation or calibration is very important.
Mesh/screen with yielding rockbolts or cables can be effective in confining the
at-risk pillar ribs. The installation of yielding elongate support such as hydraulic
props or yielding mechanical props can also be effective.
In highly critical excavations such as beltways or accessways, yielding arches can
also be considered.
Where important infrastructure passes through stiff rock structures (like dykes),
a false archway can be installed and a foamed lightweight concrete used as an infill
between the arches and the rock in the gap (typically say 50 cm). The cellular concrete
or similar infilling acts as a “shock absorber” as it can deform and absorb sudden rock
movement induced by a rockburst.

References
Agioutantis, Z., Barkley, D., Karmis, M. & Elrick, S. (2014) Development of an enhanced method-
ology for ground movement predictions due to longwall mining in the Illinois Basin. Proceedings
of the 33rd International Conference on Ground Control in Mining. Morgantown, WV.
Akande, J.M. & Saliu, M.A. (2011) Design of a powered support system in Enugu Coal Mine.
J Emerg Trends Eng Appl Sci (JETEAS). Vol. 2, No. 6, pp. 1083–1089.
182 Underground soft rock mining

Alehossein, H. & Poulsen, B.A. (2010) Stress analysis of longwall top coal caving. Int J Rock
Mech Min Sci. Vol. 47, No. 1, pp. 30–41.
Anderson, N.L., Apel, D.B., Ismail, A., Kovin, O. & Dezelic, V. (2006) Differentiating rooms
and pillars on reflection seismic profiles: A seismic investigation on two abandoned coal mines.
Highway Geophysics, NDE Conference, pp. 58–70. December.
Barasa, F. (2020) The world’s largest exporters of peat. World Facts. December 11.
Barczak, T.M. (2006) A retrospective assessment of longwall roof support with a focus on chal-
lenging accepted roof support concepts and design premises. 25th International Conference on
Ground Control in Mining. Morgantown, WV. pp. 232–243.
Bhattacharya, S. & Singh, M.M. (1985) Development of subsidence damage criteria. Engineers
International Inc. Prepared for U.S. Dept. of the Interior, Office of Surface Mining, Contract
J51120129.
Bieniawski, Z.T. (1989) Engineering Rock Mass Classifications: A Complete Manual for Engineers
and Geologists in Mining, Civil, and Petroleum Engineering. Wiley, New York.
Booth, C.J, Spande, E.D., Pattee, J.D., Miller, J.D. & Bertsch, L.P. (1998) Positive and negative
impacts of longwall mining on a limestone aquifer. Environ Geol. Vol. 34, No. 213, pp. 223–233.
70th Statistical Review of World Energy (2021) BP. www.bp.com/content/dam/bp/business-sites/
en/global/corporate/pdfs/energy-economics/statistical-review/bp-stats-review-2021-coal.
pdf ?msclkid=d416f0ddbcc811ec893b2e68db556023.
Brown, E.T., Bray, J.W., Ladanyi, B. & Hoek, E. (1983) Ground response curves for rock tunnels.
J Geotech Eng. Vol. 109, No. 1, pp. 15–39.
Bruhn, R.W. & Speck, R.C. (1986) Characteristics of subsidence over pillar extraction panels.
U.S. Bureau of Mines, Contract Report J0233920, GAI Consultants, Inc., July 1986.
Carver, L. & Rauch, H.W. (1994) Hydrogeologic effects of subsidence at a longwall mine in the
Pittsburgh Coal Mine. In: Proceeding, 13th Conference on Ground Control in Mining, edited by
Syd S. Peng, Dept. of Mining Engineering, West Virginia University.
Clark, C. & Molinda, G.M. (2007) Development and application of the Coal Mine Roof Rating
(CMRR). NIOSH (May).
Das, M.N. (1986) Influence of width to height ratio on post-failure behaviour of coal. Int J Min
Geo-Eng. Vol. 4, pp. 79–87.
Earthjustice (2009) “Waste Deep: Filling Mines with Coal Ash is Profit for Industry, but Poison
for People.” http://earthjustice.org/sites/default/files/library/reports/earthjustice_waste_deep.pdf.
Einicke, G.A. (2009) The application of smoothing within longwall mine navigation. International
Global Navigation Satellite Systems Society Symposium. Queensland, Australia. December 1–3.
Environmental Protection Agency (2016) Coal plant decommissioning, remediation and redevel-
opment. US EPA Publication No. 560-F-16-003.
Fejes, A.J. (1986) Surface subsidence over longwall panels in the Western United States. US
Bureau of Mines, Circular No. 9099.
Feng J., Peng, H., Shuai, G., Meng, X. & Lixin, L. (2016) A roof model and its application in solid
backfilling mining. Int J Min Sci Tech. Vol. 27, No. 1.
Feng, X., Hudson, G.A. & Tan, F. (eds) (2013) Rock Characterisation, Modelling and Engineering
Design Methods. CRC Press, US.
Fourie, G.A. & van Niekerk, D.J. (2001) Guidelines for the integrated planning and design of
underground coal mines. Project # 801, CSIR (November).
Garside, M. (2020) Global lignite production 1990-2018. Statista. November 30.
Gaedicke, C., Franke, D., Ladage, S., Lutz, R., Pein, M., Rebscher, D., Schauer, M., Schmidt,
S. & von Goerne (2019) BGR Energy Study. Federal Institute for Geosciences and Natural
Resources (BGR), Germany. ISBN: 978-3-9814108-7-7
Ghosh, S. S.R.K Brisbane, Australia. Mine planning software in the coal sector. SRK News Issue
#37: Focus on Coal.
Maximize Market Research (2020) Global mined anthracite market. Energy and power: Report
ID 41747. August.
Underground soft rock mining 183

Gray, R.E., Gamble, J.C., McLaren, R.J. & Rogers, D.J. (1974) State of the art of subsidence
control. Appalachian Reg. Comm., ARC-73-111-2550. December.
Hedley, D.G.F. & Grant, F. (1972) Stope and pillar design for the Elliot Lake uranium mines. Can
Inst Min Metall Bull. Vol. 65, pp. 37–44.
Hobba, W.A. (1993) Effects of underground mining and mine collapse on the hydrology of selected
basins in West Virginia. U.S. Geological Survey Water Supply Paper 2384, U.S. Department of
the Interior.
Hoek, E. & Brown, E.T. (1980) Underground Excavations in Rock. Institution of Mining and
Metallurgy, London.
Hoek, E., Kaiser, P.K. & Bawden, W.F. (1995) Support of Underground Excavations in Hard Rock.
Balkema, Rotterdam, Netherlands.
Huang, P., Spearing, A.J.S., Ju, F., Jessu, K.V., Wang, Z. & Ning, P. (2019) Control effects of five
common solid waste backfilling materials on in situ strata of gob. Energies, Vol. 12, p. 154. January.
Hunt, S.R. (1979) Characterization of subsidence profiles over room and pillar coal mines in
Illinois. Illinois Mining Institute, 86th AGM Proceedings, pp. 50–65.
Hustrulid, W.A. (1982) Underground Mining Methods Handbook. SME-AIME, New York.
Iannacchione, A.T., Mark, C., Repsher, R.C., Tuchman, R.J. & Jones, C.C. (1992) Proceedings of
the Workshop on Coal Pillar Mechanics and Design. IC9315, US Bureau of Mines.
Ingram, D.K. (1989) Surface fracture development over longwall panels in South-Central West
Virginia. US Bureau of Mines, Report 9424.
International Institute for Environment & Development (2002) Breaking New Ground:
Mining, Minerals and Sustainable Development (MMSD). Available at: http://pubs.iied.org/
9084IIED.
Jayaranjan, M.L.D., van Hullebusch, E.D. & Annachhatre, A.P. (2014) Reuse options for coal
fired power plant bottom ash and fly ash. Environmental Science and Biotechnology (April).
Jeran, P.W. & Trevits, M.A. (1995) Mining publication: Timing and duration of subsidence due to
longwall mining. US Bureau of Mines (USBM RI 9552). January.
Jiang, N., Zhao, J., Sun, X., Bai, L. & Wang, C. (2017) Use of fly-ash slurry in backfill grouting in
coal mines. Heliyon Article No.: e 00470.
Johnson, M. & Nilsson, K. (2014) Construction considerations are key in closure planning for
coal ash ponds. Power. December.
Karmis, M., Godman, G. & Hasenfus, G. (1984) Subsidence prediction techniques for longwall
and room and pillar panels in Appalachia. Proceedings of 2nd International Conference in
Stability in Underground Mining. AIME. pp. 541–553.
Kay, A. (2018) Coal 101: What is anthracite? Investing News Network. June 18.
Kennedy, C. (2013) The stages of mine design. Queens University. May (http://minewiki.
engineering.queensu.ca/).
Li, M. (ed.), Coyne, D. (2018) World coal – 2010–2050 World Energy Annual Report,
Part 4 – September).
Kim, J.S., Lee, K.S., Cho, W.J., Choi, H.J. & Cho, G.C. (2015) A comparative evaluation of stress-
strain and methods for quantitative damage assessments of brittle rock. Rock Mech Rock Eng.
Vol. 48, No. 2, pp. 495–508.
Kirsch, E.G. (1898) Die Theorie der Elastizität und die Bedürfnisse der Festigkeitslehre. Zeitschrift
des Vereines deutscher Ingenieure, 42, pp. 797–807.
Kostecki, T. & Spearing, A.J.S. (2015) Influence of backfill on coal pillar strength and floor
bearing capacity in weak floor conditions in the Illinois Basin. IJRMMS. Vol. 76, June,
pp. 56–77.
Kratzsch, H. (1983) Mining Subsidence Engineering. Springer, Berlin.
Lamé, G. (1837) Sur les surfaces isothermes dans les corps homogènes en équilibre de tempera-
ture. J Math Pures Appl. Vol. 2, pp. 147–188.
Lee, F.T. & Abel, J.F. (1983) Subsidence from underground mining: Environmental analysis and
planning considerations. Geological Survey Circular 876, US Dept. of the Interior.
184 Underground soft rock mining

Lynch, S. (2011) Mine Safety Audit Report: Metalliferous & Extractive Mines: Mine Safety
Operations Industry & Investment, New South Wales Government, Australia. February.
Mark, C. (1999) Application of coal mine roof rating (CMRR) to extended cuts. Min Eng. Vol.
51, No. 4, pp. 52–56.
Mark, C. & Agioutantis, Z. (2018) Analysis of coal pillar stability (ACPS): A new generation of
pillar design software.
Mark, C. & Gauna, M. (2015) Evaluating the risk of coal bursts in underground coal mines.
IJMST (December 15).
Mark, C. & Molinda, G.M. (2005) The Coal Mine Roof Rating (CMRR)—a decade of experi-
ence. Int J Coal Geol. Volume 64, No 1–2, pp. 85–103. October.
Michaud, D. (2017) Longwall pillar design method. 911 Metallurgist. October.
Michaud, D. (2018) Longwall mining method and design. 911 Metallurgist. May 30. www.911met-
allurgist.com/longwall-mining-methods/.
Mine Subsidence Engineering Consultants (MSEC) (2007) Introduction to longwall mining and
subsidence, www.minesubsidence.com (09.07.2015).
Missavage, R.J. (2008) Vertical shaft access. Unpublished report used for teaching. Southern
Illinois University, Carbondale.
Molinda, G., Mark, C., Bauer, E.R., Babich, D.R. & Pappas, D.M. (1998) Factors influencing
intersection stability in U.S. coal mines. Proceedings: 17th International Conference on Ground
Control in Mining. Morgantown, WV. pp. 267–275.
Monsalve, J.J., Soni, A, Ripepi, N. & Rodriguez-Marek, A. (2020) A preliminary investigation
on stochastic discrete element modeling for pillar strength determination in underground lime-
stone mines from a probabilistic risk analysis approach. 39th International Conference on
Ground Control in Mining. Canonsburg, PA.
Naganathan, S., Subramaniam, N. & Nasharuddin Bin Mustapha, K. (2012) Development of
brick using thermal power plant bottom ash and fly ash. Asian J Civil Eng. Vol. 13, No. 2,
pp. 275–287.
National Coal Board (1975) Subsidence Engineers’ Handbook. National Coal Board, Mining
Department, London.
National Institute of Occupational Safety and Health (2015) Hierarchy of controls. Workplace
safety and health topics. January.
National Resources Conservation Service (2005) Standard mine shaft and adit closing code 457.
May.
Ostrowski, W.J.S. (1972) Design considerations for modern shaft linings. CIM Transact. Vol.
LXXV, pp. 184–198.
Ozbay, M.U. & Roberts, M.K.C. (1988) Yield pillars in stope support. Proceedings of the
SANGORM Symposium in Africa, Swaziland. South African National Group on Rock
Mechanics, Johannesburg. pp. 317–326.
Palarski J. (1989) The experimental and practical results of applying backfill. In: Hassani, F.P.,
Scoble, M.J. & Yu, T.R. (eds) Innovations in Mining Backfill Technology. Balkema, Rotterdam,
pp. 33–37.
Palarski, J. (2010) Mining Technologies with Use of Fill Slurries (Thick Coal Seams). Silesian
Technical University, Poland.
Panek, L. (1956) Principles of reinforcing bedded mine roof with bolts. U.S. Bureau of Mines
RI5156.
Parizek, R.R. & Ramani, R.V. (1996) Longwall coal mines: Pre-mine monitoring and water sup-
ply replacement alternatives. Penn State Environmental Resources Research Institute, Final
Report on Legislative Initiative Program 181-90-2658. p. 194.
Peck, R.B. (1969) Advantages and limitations of the observational method in applied soil mechan-
ics. Geotechnique. Vol. 19, No. 2, pp. 171–187.
Peng, S.S. (2008) Coal Mine Ground Control, 3rd Edition, Wiley, New York, NY.
Peng, S. (2020) Longwall Mining, 3rd Edition, CRC Press. https://doi.org/10.1201/9780429260049.
Underground soft rock mining 185

Pongpanya, P., Sasaoka, T., Shimada, H., Hamanaka, A. & Wahyud, S. (2017) Numerical study
on effect of longwall mining on stability of main roadway under weak ground conditions in
Indonesia. J Geol Res Eng. pp. 93–104.
Proust, A. (1979) Thick seam winning method at the Charbonnages de France. J Mines Met Fuels.
September, pp. 285–292.
Rauch, H.W. (1985) A Summary of the Effects of Underground Coal Mines on Quantity of Ground
Water and Streamflow in the North-Central Appalachians. Eastern Mineral Law Foundation,
Sheraton Hotel at Station Square, Pittsburgh, PA.
Roseke, B. (2017) Steps in project scheduling. Project Engineer. (October).
Salamon, M.D.G. & Munro, A.H. (1967) A study of the strength of coal pillars. J South Afr Inst
Min Metallurgy. Vol. 68, pp. 55–67.
Seedsman, R. (2002) The strength of the pillar-floor system. Proceedings of 12th Coal Operators
Conference, University of Wollongong, New South Wales, Australia. pp. 24–31.
Singh, M. (1986) Mine subsidence. SME. September.
Singh, K.B. & Dhar, B.B. (1997) Sinkhole subsidence due to mining. Geotech Geol Eng. Vol. 15,
pp. 327–341.
Soliman, M.M. (1998) Environmental Hydrogeology. CRC Press LLC, US. pp. 81–101.
Solymos, A. (1983) The development of mining thick seams of coal. Mining Engineer. March.
Spearing, A.J.S. (1988) Hard rock pillar design. Internal report for AAC Gold Division,
Johannesburg, South Africa.
Spearing, A.J.S. (1994) An overview of coal mining in Poland. J S Afr Inst Min Metallurgy.
August.
Spearing, A.J.S. (2008) Free standing adjustable steel based supports in coal mines in the
USA. 27th International Conference on Ground Control in Mining. Morgantown, WV.
pp. 366–373.
Spearing, A.J.S., Benton, D., Kostecki, T. & Hirschi, J.C. (2013a) The potential of coal washing
plant waste as a backfill in room-and-pillar coal mines. Proceedings of the 32nd International
Ground Control Conference. pp. 68–75.
Spearing, A.J.S., Tadolini, S.C. & Kostecki, T.R. (2013b) The design of entry support in coal mine
room and pillar mines in the USA to combat rock related falls of ground. 23rd World Mining
Conference, Montreal, Canada. August 11–15.
Speers, C.R. & Spearing, A.J.S. (1996) The design of tunnel support in deep hard rock mines
under quasistatic conditions. J S Afr Inst Min Metallurgy. Vol. 96, No. 6, p. 295.
Sun, H., Ma, L., Liu, W., Spearing, A.J.S., Han, J. & Fu, Y. (2021) The response mechanism of
acoustic and thermal effect when stress causes rock damage. Applied Acoustics (April 2).
Surawar, S (2014) Organic rocks and fossil fuels. Slideshare. March. www.slideshare.net/
sanathsurawar/organic-rocks-and-fossil-fuels.
Terzaghi, K. (1943) Theoretical Soil Mechanics. John Wiley & Sons. ISBN 0-471-85305-4. https://
doi.org/10.1002/9780470172766.
Thi, N.N., Thinh Hong, T.P. & Truong, S.B. (2019) Power plants in Vietnam as partial replace-
ment of aggregates in concrete pavement. Hindawi J Eng (October 17).
Timoshenko, S. & Woinowsky-Krieger, S. (1959) Theory of Plates and Shells. McGraw–Hill, New
York.
University of Wollongong, Mining Science & Technology (2020) Chock shield supports. www.
miningst.com/longwall-mining/equipment/powered-supports/chock-shield-supports.
University of Wollongong, Mining Science & Technology (2020) Type of pillars. www.miningst.
com/longwall-mining/ground-control/type-of-pillars/.
US Corps of Engineers (1980) Engineering and design: Rock reinforcement. US Corps of
Engineering. Engineer Manual 1110-1-2907. February 15.
Wagner, D. (1980) Pillar design in coal mines. J S Afr Inst Min Metallurgy. pp. 37–45. January.
Whittaker, B.N. & Reddish, D.J. (1989) Subsidence: Occurrence, Prediction and Control.
Department of Mining Engineering, The University of Nottingham, UK.
186 Underground soft rock mining

Zenc, M. (1969) Comparison of Bals’ and Knothe’s methods of calculating surface movements
due to underground mining. Int J Rock Mech Min, Sei 6.
Zhou, Y., Zhang, D., Fan, G., Zhang, S. & Zhang, S. (2020) Feasibility study on fully mechanized
large mining height long wall top-coal caving mining in ultra-thick (20–30 m), parting-rich coal
seams: A case study of the Laosangou mining field in China. Energy Sources.
Zhang, J., Li, M., Taheri, A., Zhang, W., Wu, Z. & Song, W. (2019) Properties and application of
backfill materials in coal mines in China. Minerals. Vol. 9, No. 1, p. 53.
Zhao, Y.X. & Jiang, Y.D. (2010) Acoustic emission and thermal infrared precursors associated
with bump-prone coal failure. Int J Coal Geol. Vol. 83, No. 1, pp. 11–20.
Chapter 3

Underground hard rock (metal/


non-metal) mining

3.1 Introduction to hard rock mining

3.1.1 Typical hard rock geology


Unlike soft rock mining (typically coal), hard rock deposits occur in not only sedi-
mentary rock formations but also in igneous and metamorphic rocks. Rather than
mining a coal or salt seam, in hard rock mining, we are exploring for and then mining
an orebody. Ore is commonly defined as a mineral(s) that can be exploited safely and
at a profit. Within the boundaries of an orebody, there are often low-value or no-value
minerals which are referred to as internal waste or gangue materials.
Frequently, the orebody does not outcrop and could have considerable cover (overbur-
den) above it; therefore, finding such orebodies can be technically challenging and costly.

3.1.2 Hard rock orebody formation and classification


According to Lacy (2015), typically orebody geological formation processes include:

• Weathering and general erosion – This can break down the rocks into miner-
als which may become concentrated as either weathering-resistant minerals left
behind, or in the mobile minerals being removed and transported in solution or
carried as solid particles and then concentrated during the erosional processes.
• Volcanic processes – These can result in concentrations through either crystalli-
zation and differential gravity settling, concentration in end phases of crystalliza-
tion, evolved fluids or as a consequence of heat energy introduced by the intrusives
(rocks formed after lava slowly cools and then crystallizes).
• Sedimentation – This process can result in gravity separation and concentration
within clastic (small weathered mineral particles) sediments in streams, lakes and
by ocean shore currents. As pore-waters are expelled from these compacting sed-
iments, they may selectively leach, carry and preferentially deposit specific min-
erals or elements.
• Tectonic or metamorphic processes – These processes may result in the break-
down and transformation of rock minerals and concentration of certain elements
by expelled fluids or diffusion.

Lacy (2015) further suggests that mineral deposits can be defined by considering the
geological environment in which they are formed and found. This is outlined below.

DOI: 10.1201/9781003185680-3
188 Underground hard rock mining

3.1.2.1 Igneous intrusions


These typically include:

• Iron
• Ferro-alloy metals (e.g., chromium, tungsten, nickel, manganese and tin)
• Non-ferrous base metals (e.g., copper, lead, zinc and tin)
• Precious metals (silver and gold)
• Nuclear elements (e.g., uranium, strontium and thallium)
• Magnesite (a magnesium carbonate mineral)
• Barite (a barium sulphate mineral)

3.1.2.2 Sedimentary concentration by weathering and/or precipitation


This includes:

• Iron oxides
• Manganese oxides
• Bauxite
• Uranium
• Vanadium
• Copper
• Nickel

3.1.2.3 Sea water extraction


This includes:

• Salt
• Magnesium

3.1.2.4 Placer deposits (beaches and streams)


These include:

• Gold
• Cassiterite (tin)
• Mercury
• Zirconium
• Silica sand
• Rutile (titanium dioxide) and ilmenite (an oxide of iron and titanium)
• Monazite (thorium and some rare earths)

3.1.2.5 Pegmatite deposits (igneous deposits caused by slow


crystallization at high temperatures and pressures)
These include:

• Feldspar (aluminium tectosilicate minerals and the most common rock in the
earth’s crust)
Underground hard rock mining 189

• Quartz
• Rare earth minerals
• Lithium minerals (most common form is spodumene)
• Beryllium

3.1.2.6 Evaporite deposits


These deposits include:

• Potash (mainly potassium chloride)


• Sodium carbonate
• Sodium sulphate
• Salt
• Borax (borates)
• Gypsum
• Lithium (lithium salts)

3.1.2.7 Chemically deposited sedimentary deposits


These deposits include:

• Sulphur
• Phosphate
• Strontium minerals (mainly celestite and strontianite)
• Barite (barium sulphate)
• Limestone
• Dolomite
• Magnesite
• Chalk

3.1.2.8 Other sedimentary deposits


These include:

• Diatomite (rock formed from diatomaceous earth)


• Clay
• Shale
• Sandstone
• Quartzite (silica sand)

3.1.2.9 Volcanic deposits


These include:

• Bentonite (a swelling clay of mainly montmorillonite)


• Perlite (an amorphous very low-density volcanic glass)
• Diamond-bearing rock (typically kimberlites)
190 Underground hard rock mining

3.1.2.10 Metamorphic deposits


This is a broad category and includes:

• Graphite
• Mica
• Asbestos
• Talc
• Vermiculite (a phyllosilicate mineral)
• Pyrophyllite (a phyllosilicate mineral of mainly aluminium silicate hydroxide)
• Slate (metamorphized shale)
• Sillimanites or fibrolite minerals (including andalusite)
• Magnesite (magnesium carbonate)

During early prospecting and exploration in the field, geologists often look for specific
indicators such as:

• Rust (Gossan) – Rust-coloured rocks or soil are one of the best indications of
mineralization. Rust is caused by iron, which may indicate the presence of other
metallic minerals. A variety of rust-coloured minerals exist. Copper sulphide is
identified by green or blue stain, nickel sulphide by a pale green stain and ura-
nium minerals may weather to bright yellow, orange or green stains.
• Quartz – Most sulphide mineral deposits are associated with quartz veins. Unless
visible, gold and silver cannot be definitely determined without an assay test, but
other metals such as copper, lead and zinc may be identified without an assay.
• Dykes – These are long thin bodies of igneous rock that, while in a molten state,
flowed into cracks in older rocks. They stand out from the rock they flowed into
and frequently contain or are associated with valuable minerals.
• Shear zones – A shear zone is a place of weakness or a break in the earth’s crust
through which mineralized solutions may have been channelled. From the air,
they appear as long depressions or lines. Since they often contain quartz veins,
they are a likely location of metallic minerals.

3.1.3 Hard rock exploration

3.1.3.1 Introduction to exploration


Many years ago, and it still remains true today, Harry Parker of AMEX stated the
following as it pertains to exploration:

Trust no-one
Assume nothing
Check Everything!

Prospecting is the early stage of exploration and is the physical search for minerals,
fossils, precious metals or mineral specimens, and was also known as fossicking.
Prospecting still requires significant physical labour and often involves covering large
remote areas of potential interest using four-wheel drive vehicles, all-terrain vehicles
Underground hard rock mining 191

(ATVs), horses or just on foot. It can involve panning, outcrop investigation, sampling
with typically a geological hammer and looking for signs of significant mineralization.
A prospector must also first stake claims and follow all regional and local regulations
and work closely with the community where the prospecting is occurring.
Mineral exploration is the whole investigation undertaken by geologists working
alone or more commonly from companies or similar organizations to try and find new
and potentially commercially viable mineral deposits. Mineral exploration should be
thorough, intensive, organized and professionally undertaken with all samples and
data being carefully recorded, stored and analyzed.

3.1.3.2 The Stages of Exploration


The stages in exploration are:

• Literature review, which involves studying all Geological Survey and Government
or State information if available in a particular country and all related previ-
ous exploration and mining in the areas of interest. After this stage, exploration
licenses in the areas of interest should have been registered, approved and permit-
ted by the relevant regulatory authority.
• Field surveys involves focusing on any favourable sites identified in the literature
review. Geologists then conduct field surveys to determine the mineral potential
in a target area. To start, they take samples from any mineralized outcrops in the
area. These are rocks that are part of a formation, and generally not loose boul-
ders (which may have been transported from another area by rain, for example,
and that are visible from the surface. This sampling and analysis (assaying) will
indicate the type, nature and characteristics of the formation, including the min-
erals present. Sometimes a few outcrops are enough to obtain a first rough pic-
ture of the mineral deposit and geology of the area. After the samples are tested,
labelled and recorded with a brief description of their location and geological
features, chemical tests/assays are conducted. All the information is plotted on a
map showing the precise locations of sample sites and findings.
• Geotechnical and other studies/surveys are undertaken, which can include: pho-
togeological surveys, geophysical surveys, geochemical surveys, geobotanical
surveys, radar scanning, infrared scanning and physical property testing. These
include:
• Airborne studies, which can be relatively quick and cost effective and can also
be used to map infrastructure (such as roads, fences, villages, etc.).
• Checking for variations in gravity, magnetic anomalies, radioactivity and
rock resistivity/conductivity.
• Assaying (for example, stream sediments).
• A logically established grid sampling system is frequently used.
• If the area is covered by alluvial soil (transported material), shallow holes
may need to be drilled to reach bedrock.
• Diamond drilling programs are essential as any geophysical survey needs precise
secondary exploration by drilling to confirm the orebody delineation and more pre-
cisely identify the minerals present and obtain a grade indication. If the mineral-
ization outcrops or nearly outcrops, then trenching into the deposit is also useful.
192 Underground hard rock mining

• Market and feasibility studies


• A feasibility study evaluates the size and depth of the orebody and recom-
mends mining methods to safely and cost effectively extract it. These studies
also establish rates of production. Investigative tools could include close-
spaced diamond drilling and perhaps underground investigation.
• Feasibility and market studies are required by major investors (e.g., banks)
before they will commit funds to the project. It is at this make-or-break stage
in the exploration process that decisions are made to pursue development,
temporarily shelve the project until it becomes economically more attractive
or abandon it. Frequently, what had been considered an exciting property
initially is abandoned when the accumulated evidence shows that the deposit
is too small or insignificant to mine profitably at current commodity prices.
• As with other exploration stages, positive market and feasibility studies
enhance the value of the property and related mineral rights. Negative results,
however, could mean that all the time and money invested in the exploration
to date lead to a decision not to proceed. This underlines the importance of
strong positive results at each stage of exploration to warrant taking the next
step and spending more money.
These previous activities aim at the following outcomes:

• Mineral of interest (generally before the Literature Review)


• Area selection (after the Literature Review)
• Target definition during Geotechnical Studies and Drilling and involving:
• Preliminary investigations
• Geological mapping
• Geophysical and/or geochemical testing of the surface and subsurface
• Drilling and/or trenching
• Resource evaluation after most of the Geotechnical Studies and Drilling and involving:
• A first quantification of the grade and tonnage of the orebody, achieved by
drilling on a regular grid pattern
• Drilling additional holes in “areas of interest” (e.g., major fault zones, etc.)
• Depending on the orebody (e.g., sulphides), the holes could also be used for
detailed geophysical probing
• Reserve definition (after more detailed diamond drilling), which converts a min-
eral resource into an ore reserve and involves:
• More intensive and technical evaluation than resource evaluation aimed at
statistically grading the orebody and geometry
• Metallurgical testing and rock mechanics assessments
• The aim is to produce a full bankable feasibility study proving the economic
viability
• Environmental Impact Assessment (EIA)
• Before construction of a mine can start, an environmental assessment is usu-
ally required by law.
Underground hard rock mining 193

• Environmental assessments are rigorous, technical and have their own set of
criteria. Technical experts should be brought in to conduct them properly.
There are environmental consulting firms familiar with all aspects of EIAs in
most countries.
• All interested parties, especially the local community, should be involved,
and the whole process needs to be thorough and transparent.
• The costs of environmental assessments vary depending on the size of a pro-
ject but are normally less than 1% of the total cost of the project.
• Trial/full extraction
• Reclamation

Exploration areas are separated into two main types:

• Greenfields
• These are the most difficult as there is no existing exploration or related min-
ing in the area.
• They have higher risk, lower success rate but higher rewards.
• Brownfields
• These are generally near existing related mines or well-known geologically
detailed areas.

3.1.3.3 More details on exploration methods


This section gives more details of the possible testing and methods used to help define
and prove an orebody (ignoring diamond drilling which is discussed later). Typical
exploration methods, after initial prospecting and exploration activities and investi-
gations, include:

• Photogeological surveys
• Geophysical surveys
• Geochemical surveys
• Geobotanical surveys
• Radar scanning (such as ground penetrating radar)
• Infrared scanning

Photogeological surveys are undertaken using drones, aircraft or satellites and involve
taking many high-resolution photographs to determine the geological (and hydrogeo-
logical) structure of the surface. They can help indicate the possible sub-surface geo-
logical structure too.
There are several different geophysical surveys that can be undertaken, including:

• Magnetic surveys – These surveys identify local variations (anomalies) in the


earth’s magnetic field. They are particularly useful in locating magnetite, ilmenite
and pyrrhotite. Magnetic logging in drill holes can be used to direct other explo-
ration drilling.
194 Underground hard rock mining

• Electromagnetic surveys – The theory behind these tests is that a primary alter-
nating current (AC) induces a secondary current in the presence of a conductor.
This method is therefore typically useful for copper and lead sulphide ores, mag-
netite, pyrite, certain manganese ores and graphite.
• Electrical surveys – These surveys measure the natural electrical flow in the rock
(galvanic currents). They can be used to find shallow deposits, to map geological
features and to locate the water table. In exploration, they can be used to further
differentiate the zones located by electromagnetic surveys.
• Gravimetric surveys – These use the small changes in gravity caused by certain
rocks and rock formations. They can help locate major faults, synclines and anti-
clines, salt domes, oil-bearing formations and certain high-density orebodies
(e.g., iron ore).
• Radioactivity surveys – These surveys identify increases in the local natural radi-
ation level in the ground. They are specifically used to find uranium, thorium or
other minerals associated with radioactive materials.
• Seismic surveys – The basis of this method is that different strata (rock layers)
have different sound wave velocities. These sound waves can be manmade by huge
surface hammers, explosives or a mechanical vibrator. They can be used to deter-
mine oil-bearing strata, surfaces of ore-bearing strata, etc.

Geochemical surveys are done by taking samples from soils, stream and river sedi-
ments to try and identify a geological anomaly that could indicate the presence of an
orebody.
Geobotanical surveys involve the identification of certain plants (often flowering)
that are known to grow in locations that are richer in certain minerals or rocks. It is
most commonly used to identify the presence of uranium.
Radar scanning is used to determine geology and hydrology near the surface (to
a depth of typically 30m). It is useful to determine soil depths, the presence of major
fractures, hydrology and different rock strata.
Infrared scanning can identify different geological structures and rock strata.

3.1.3.4 Diamond (core) drilling


Any geophysical survey needs a secondary exploration phase undertaken using core
drilling to confirm the orebody delineation and more precisely identify the minerals
present and obtain a grade indication.
If the mineralization outcrops or nearly outcrops, then trenching into the deposit is
also useful. Diamond drilling gives critical information needed to establish a viable
orebody and can be relatively costly.
The diamond drill rod produces a cylinder of the rock strata or orebody that
can be removed and evaluated and tested (typically for rock properties, mineral
identification, geology and grade). Any core drilling program must be inten-
sive enough to adequately delineate the orebody and obtain a reliable grade
indication.
The components of a diamond drill rig are shown in Figure 3.1.
The diamond drilling bit consists usually of industrial diamonds in a strong matrix
to be able to cut the rock. A drill bit with wedges that are used to keep the drilled core
in place as it is being removed from the hole is shown in Figure 3.2.
Underground hard rock mining 195

Figure 3.1 Components of a diamond drill rig (Rodriguez, 2017)

Figure 3.2 A diamond drill bit


Reproduced with kind permission from Tae Sung Co.
196 Underground hard rock mining

Long-term secure core storage is essential especially if there is a subsequent need


for re-evaluation. Cores are very important as they represent the key exploration tool
(giving geometry and structure, grade and a geotechnical assessment).

3.1.3.5 Rock sampling


Sampling is the process of estimating the mineral content and other properties of an
ore deposit by averaging the results of a number of small representative portions (sam-
ples) taken from the specific rockmass. The nature and consistency of the orebody
must be considered as it has a huge effect on the sampling needs.
Sources of samples for new mining projects could be obtained from:

• Orebody outcrops
• Drill core
• Blast hole cuttings
• Trenches
• Exploration adits, shafts or declines
• Bulk sampling for metallurgical testing

The types of samples that can be used include:

• Bulk samples
• Groove samples
• Chip samples
• Drill hole (preferably diamond-drilled) samples
• Grab samples

Sampling is a critical part of exploration and mining to determine the size (bounda-
ries) and grade (value) distribution. Sampling is also critical to relate the head grade
of the run of mine ore on the surface and the recovery after processing to control and
improve the extraction recovery.
Sampling is subject to both random and bias errors. A high level of sampling impar-
tiality is needed to try and reduce at least the bias errors.
Other sampling can also be undertaken during mining to determine the following:

• The air quantity in any working place


• The dust and certain gas levels (such as methane or radon)
• The water quality throughout the mine and especially at discharge out of the mine
property
• Alcohol and drug testing (sampling) of personnel to help ensure on-the-job safety

3.1.4 Hard rock exploitation


Hard rock deposits can be mined using any of the common mining methods (or com-
binations), mainly depending on:

• Orebody dimensions
• Orebody dip (gravity)
• Depth range and the stress tensor
Underground hard rock mining 197

• Orebody properties
• Commodity value
• Orebody grade and distribution
• Host rock properties
• Regional geology and hydrology
• Local geology
• Socio-economic considerations
• Location and infrastructure
• Relevant regulations

Common hard rock mining methods include:

• Open pit mining


• Room and pillar and drift mining (usually with backfill)
• Longwall mining (but not similar to longwall coal mining; it only refers to the
overall mining faces in hard rock tabular mines)
• Shrinkage mining
• Cut and fill
• Undercut and fill
• Longhole open stoping
• Vertical crater retreat (VCR)
• Sub-level caving
• Block caving

Figure 3.3 shows the outline of the mining method selection.

3.2 Basic hard rock mine design and planning

3.2.1 Hard rock mine design principles


Relating laboratory rock tests (such as unconfined compression and tensile tests, or
triaxial compressive tests) to the actual performance of the rockmass as a whole is the
main design challenge in any mining operation, whether surface or underground and
whether in hard or soft rock.
Rockmass classification systems are therefore needed because a rockmass does not
behave as a small core sample tested in the laboratory. In the rockmass, the effec-
tive strength is significantly reduced by water and geological features such as bedding
planes, faults and joint sets. Classification systems try and account for this, and in a
greenfield development (i.e., an area in which no similar type of mines exist), this is all
the available information in most cases.
The objectives and sequence of any rockmass classification system are to:

• Determine all the significant factors that degrade the rockmass as a whole
• Define the rockmass of interest into individual groups all having the same approx-
imate mechanical properties
• Establish the qualitative properties of each individual group and assign a quanti-
tative number or percentage to each
• Use this information for a first pass of a mine layout and support design
198 Underground hard rock mining

Figure 3.3 The mining method selection


Underground hard rock mining 199

The above generally requires close involvement of mainly geologists and mining engineers.
There are many hard rock rockmass rating systems available, all empirically based,
but only the following will be discussed:

• Rock Quality Designation (RQD)


• Q system
• Rock Mass Rating (RMR)
• Mine Rock Mass Rating (MRMR)

Rockmass classification or rating methods must not be used in isolation and should
be one tool in the mine method selection and underground excavation design process.
One of the main reasons why any rockmass rating system should not be used in isola-
tion is that the actual geology of the new mine may be significantly different from that
used in the development of any such empirically based rockmass rating system. The
methods (except RQD) are also interpretive and can vary significantly depending on
the individual applying them and their unique experience.
Geomechanics or rock mechanics using computer-based analyses are very useful mine
design tools and should always be considered, with care to validate or calibrate them.
The goal of all these methods is to be able to design safe, stable and economic mine
infrastructure and production units (whether stopes, panels, longwalls or benches)
based mainly on the local geological, hydrological data and laboratory tests (usually
obtained from cores).

3.2.1.1 Rock Quality Designation (RQD)


This system was developed in 1964 by D. Deere. It is based on measuring the core
recovery percentage, which incorporates only pieces that are greater than 100 mm
in length. In this respect, pieces of core that are not hard and sound should not be
counted though they may be 100 mm or longer in length. The optimal core diameter
for RQD is 47.5 mm.
RQD has value in estimating the support required for excavations. This quantita-
tive index is used specifically to identify low-quality rock zones. RQD is used now as
a standard parameter in drill core logging and forms a basic element value of all the
main mass classification systems.
RQD is defined as:

RQD = [(Sum of length + 10cm pieces)/(Total length of core)] *100%

This is then used to categorize the specific rock stratum broadly as follows in Table 3.1:

Table 3.1 The relationship of RQD to rock quality

Rock Quality Measured RQD (%)

Very poor 0–25


Poor 25–50
Fair 50–75
Good 75–90
Excellent 90–100
200 Underground hard rock mining

Figure 3.4 An RQD example (Deere & Deere, 1988)

The main caution is that any “new” fractures caused by the diamond drilling itself
(e.g., from poor operation or worn drill bits) should not be included in defining the
core lengths if possible. It is not too serious if such “man or machine fractures” are
included because they just introduce more conservatism into the design, which makes
it safer and not more hazardous.
The advantage of RQD is that the value is not prone to interpretation or bias.
Like all empirical systems, however, it has limitations. For example, if a core con-
sists wholly of pieces between say 9 and 10 cm, it will have an RQD near zero, and if
another consists wholly of pieces between 10 and 11 cm, the RQD will be near 100%
and yet the two rock strata would be relatively similar in reality. The effects of water
on the strength are also ignored.
Figure 3.4 gives an RQD example (Deere & Deere, 1988).

3.2.1.2 Q System
The Q System (after Barton, 1976) further refines and expands on the RQD system and
tries to account for the presence of joints, water and the stress state, based on over 200
underground (mainly civil tunnel) case studies in Scandinavia:

Q = (RQD/Jn)(Jr/Ja)(Jw/SRF)
Underground hard rock mining 201

Where:
• Q is the overall quality out of 100%
• RQD is the rock quality designation
• Jn is a joint set number
• Jr is a roughness number
• Ja is a joint alteration number
• Jw is a joint water reduction factor
• SRF is the stress reduction factor
The terms making up the value Q can be explained as follows:
• (RQD/Jn) is an indication of block size in the rock mass
• (Jr/Ja) is an indication of the shear strength between the blocks in the rock mass
• (Jw/SRF) is a function of the actual stress acting on the rock mass
The main shortcomings of this system are that they are focused on the specific com-
plex geology in tunnels in Norway and don’t seem to adequately take account of joint
direction and joint continuity.
Factors making up the Q System:
THE RQD

How to calculate RQD is given earlier and the ranking is in Table 3.2.
THE JOINT SET NUMBER

Table 3.2 Joint Set Number (Jn)

A-Massive, no or few joints 0.5–1.0 Notes


B-One joint set 2
C-One joint set plus random 3
D-Two joint sets 4
E-Two joint sets plus random 6
F-Three joint sets 9 Note: For intense, use (3.0 × Jn)
G-Three joint sets plus random 12 Note: For portals, use (2.0 × Jn)
H-Four or more joint sets, 15
random, heavily jointed, “sugar
cube”, etc.
J-Crushed rock, earth-like 20

An example of a jointed rockmass is given in Figure 3.5 (after Kim et al., 2007).

Figure 3.5 A rockmass with joints sets (Kim et al., 2007)


202 Underground hard rock mining

THE JOINT ROUGHNESS NUMBER

Table 3.3 Joint Roughness Number (Jr)

Factor Notes
(a) Rock-wall contact
(b) Rock-wall contact before
10-cm shear
A-Discontinuous joints 4
B-Rough or irregular, undulating 3 Note: Descriptions refer to small-
C-Smooth, undulating 2 scale features and intermediate-
D-Slickensided, undulating 1.5 scale features, in that order
E-Rough or irregular, planar 1.5 Note: Add 1.0, if the mean spacing
F-Smooth, planar 1.0 of the relevant joint set is greater
G-Slickensided, planar 0.5 than 3m
(c) No rock-wall contact when
sheared
H-Zone containing clay minerals 1.0 Note: Jr = 0.5 can be used for
thick enough to prevent planar slickensided joints having
rock-wall contact lineations provided the lineations,
J-Sandy, gravelly or crushed zone 1.0 are orientated for minimum
thick enough to prevent strength
rock-wall contact

Joint roughness is quite subjective and difficult to determine. Figure 3.6 shows a
“rough/uneven” joint and a Barton Comb used to measure the joint roughness.

Figure 3.6 Joint roughness being measured using a Barton Comb (Morelli, 2014)
Underground hard rock mining 203

JOINT ALTERATION NUMBER

Table 3.4 Joint Alteration Number (Ja)

Factor (φ friction angle)


(a) Rock-wall contact
A-Tightly healed, hard, non-softening, impermeable filling, i.e., 0.75 (-)
quartz or epidote
B-Unaltered joint wall, surface staining only 1.0 (25–35o)
C-Slightly altered non-softening mineral coatings, sandy particles, 2.0 (25–30o)
clay-free disintegrated rock, etc.
D-Silty or sandy-clay coatings, small clay fraction (non-soft) 3.0 (20–25o)
E-Softening or low-friction clay mineral coatings, i.e., kaolinite 4.0 (8–16o)
or mica. Also, chlorite, talc, gypsum, graphite, etc., and small
quantities of swelling clays
(b) Rock-wall contact before 10-cm shear
F-Sandy particles, clay-free disintegrated rock, etc. 4.0 (25–30o)
G-Strongly over-consolidated non-softening clay material fillings 6.0 (16–24o)
(continuous but <5 mm thickness)
H-Medium or low over-consolidation, softening, clay mineral 8.0 (12–16o)
fillings (continuous but <5 mm thickness)
J-Swelling clay fillings, i.e., montmorillonite (continuous, 8–12 (6–12o)
but <5 mm thickness).Value Ja depends on percentage of
swelling clay-sized particles and access to water, etc.
(c) No rock-wall contact when sheared
K-Zones or bands of disintegrated or crushes rock 6, 8 (6–24o)
or 8–12
L-and clay (see G, H, J above for description of clay condition) 5.0 (-)
M-Zones or bands of silty or sand clay, small clay fraction
(non-softening)
N-Thick, continuous zones or bands of clay (see G, H, J above 10, 13 (6–24o)
for description of clay condition) or 13–20

JOINT WATER REDUCTION FACTOR

Table 3.5 Joint Water Reduction Factor (Jw)

Water condition Factor Approx. water pressure, kPa/cm2

A-Dry excavations or minor inflow, i.e., 1.0 <1.0


<5 litres/min locally
B-Medium inflow or pressure, occasional 0.66 1–2.5
outwash of joint fillings
C-Large inflow of high pressure in competent 0.5 2.5–10
rock with unfilled joints Note: Factors C–F are crude
estimates. Increase
D-Large inflow of high pressure, considerable Jw if drainage measures are installed
outwash of joint fillings
E-Exceptionally high inflow or water pressure 0.2–0.1 >10
at blasting, decaying with time Note: Special problems caused
by ice formation are not
considered
F-Exceptionally high inflow or water pressure 0.1–0.05 >10
continuing without noticeable decay
204 Underground hard rock mining

Figure 3.7 The effects of water on rock (a rock Island in the South China Sea)
Photo taken by Sam Spearing

Figure 3.7 shows how water can seep through porous rocks and leach material out
of the rockmass (forming the white stalactites) and weaken it. In addition, the tidal
motion of the sea has undermined the rockmass, and in geological time, it is likely to
totally collapse as the undermining continues.
STRESS REDUCTION FACTOR

The following from Stillborg (1986) is relevant and should be considered:

“When borehole core is unavailable, the RQD can be estimated from the number
of joints per unit volume, in which the number of joints per metre for each joint
set are added. A simple relation can be used to convert this number to RQD for
the case of clay free rock masses:

RQD ≈ 115 − 3.3 Jv

Where:

Jv = total number of joints per m3 (RQD = 100 for Jv < 4.5).


Jn, which represents the number of joint sets, will often be affected by foliation,
schistosity, slaty cleavage or bedding. If strongly developed, these parallel “joints”
should obviously be counted as a complete joint set. However, if there are few “joints”
visible, or only occasional breaks in the core due to these features, then it will be more
appropriate to count them as “random joints” when evaluating the Jn parameter.
The parameters Jr and Ja (representing the rock shear strength) should be
related to the weakest significant joint set or clay-filled discontinuity in the given
zone. However, if the joint set or discontinuity with the minimum value of (Jr/Ja)
Underground hard rock mining 205

Table 3.6 Stress Reduction Factor (SRF)

Weakness zones intersecting excavation, Factor Note: Reduce these values of SRF by
which may cause loosening of rock 25–50% if the relevant shear zones
mass when tunnels are excavated only influence but do not intersect
A-Multiple occurrences of weakness 10
zones containing clay or chemically
disintegrated rock, very loose
surrounding rock (any depth)
B-Single weakness zones containing clay 5
or chemically disintegrated rock (depth
of excavation <50m
C-Single weakness zones containing clay 2.5
or chemically disintegrated rock (depth
of >50m)
D-Multiple shear zones in competent 7.5
rock (clay-free), loose surrounding rock
(any depth)
E-Single shear zone in competent rock 5.0
(clay-free) (depth of excavation <50m)
F-Single shear zone in competent rock 2.5
(clay-free) (depth of excavation >50m)
G-Loose open joints, heavily jointed or 5.0
“sugar cube”, etc. (any depth)
(b) Competent rock, rock stress problems σc/σ1 σt/σ1 (SRF)
H-Low stress, near surface >200 >13 2.5 Note: For strongly anisotropic virgin
J-Medium stress 200-10 13-0.66 1.0 stress field (if measured): When 5 ≤
K-High stress, very tight structure 10-5 0.66-0.33 0.5-2 σ1/σ3 < 10, reduce σc and σt to 0.8 σc
(usually favourable to stability, may be and 0.8 σt. When σ1/σ3 > 10, reduce
unfavourable for wall stability) σc and σt, 0.6 σc and 0.6 σt, where σc
L-Mild rockburst (massive rock) 5-2.5 0.33-0.16 5-10 = unconfined compression strength
M-Heavy rockburst (massive rock) 0.16 10-20 and σt = tensile strength (point load),
Squeezing rock: plastic flow of and σ1 and σ3 are the major and minor
incompetent rock under the influence principal stresses
of high rock pressure
N-Mild squeezing rock pressure 5–10 Note: Few case records available where
O-Heavy squeezing rock pressure 10–20 depth of crown below surface is less
Swelling rock: chemical swelling activity than span width. Suggest SRF increase
depending on the presence of water from 2.5 to 5 for such cases (see H)
P-Mild swelling rock pressure 5–10
R-Heavy swelling rock pressure 10–15

is favourably orientated for stability, then its higher value of Jr/Ja should be used
when evaluating Q. The value of Jr/Ja used should in fact relate to the surface most
likely to allow failure to initiate.
When a rock mass contains clay, the factor SRF appropriate to loosening
loads should be evaluated. In such cases, the strength of the intact rock is
of less relevance. However, when jointing is minimal and clay is completely
absent, the strength of the intact rock is of little interest. However, when joint-
ing is minimal and clay is completely absent, the strength of the intact rock
may become the weakest link, and the stability will then depend on the ratio of
the rock-stress/rock-strength. A strongly anisotropic stress field is unfavoura-
ble for the stability and is roughly accounted for in Table 3.6 for stress reduc-
tion factor evaluation.”
206 Underground hard rock mining

Table 3.7 Excavation Support Ratio (ESR) for different excavations

Ranking Excavation category ESR

A Temporary mine openings 3–5


B Permanent mine openings, water 1.6
tunnels for low pressures, pilot
tunnels, drifts and headings for large
excavations
C Storage rooms, water treatment 1.3
plants, minor road and rail tunnels,
surge chambers and access tunnels
D Power stations, major road and rail 1.0
tunnels, civil defence chambers,
portals and excavation intersections
E Nuclear power stations, railway 0.8
stations, sports and public facilities
and factories

The compressive and tensile strengths (σc and σt) of the intact rock should be eval-
uated in the saturated conditions if this is appropriate to present or projected future
in-situ conditions. A very conservative estimate of strength should be made for those
rocks that deteriorate when exposed to moist or saturated conditions.
In order to relate the calculated Q value to the support requirement underground,
Barton et al. (1974) developed a further concept called Equivalent Dimension (De):

De = (Excavation span, diameter or height)/(Excavation Support Ratio)

The Excavation Support Ratio (ESR) is related to the use of the excavation and the
time the excavation is required to be stable for.
Barton (1976) gives the following values (Table 3.7) for the ESR.
Figure 3.8 (Barton et al., 1974) gives a derived empirically based support require-
ment guide.

3.2.1.3 Rock Mass Rating (RMR)


This geomechanics system was developed by Bieniawski (1973) and has been exten-
sively adjusted and expanded since then.
This system uses six parameters to define the rockmass:

• RQD
• Unconfined compressive strength (UCS) of a particular rockmass
• Joint spacing
• Joint condition
• Joint orientation
• Hydrology

These parameters can be estimated from core samples or rockmasses with similar
geological features that are exposed by mining.
The following tables are taken from Bieniawski (1989).
Table 3.8 RMR Classification parameters (Bieniawski, 1989)

A. CLASSIFICATION PARAMETERS AND THEIR RATINGS

Parameter Range of values (ratings)

1 Strength of intact Point-load strength >10 MPa 4–10 MPa 2–4 MPa 1–2 MPa For this low range, uniaxial
rock material index compressive strength is preferred
Uniaxial >250 MPa 100–250 MPa 50–100 MPa 25–50 MPa 5–25 MPa 1–5 MPa <1 MPa
compressive
strength
Rating 15 12 7 4 2 1 0
2 Drill core quality RQD 90–100% 75–90% 50–75% 25–50% <25%
Rating 20 17 13 8 5
3 Spacing of discontinuities >2m 0.6–2m 200–600 mm 60–200 mm <60 mm
Rating 20 15 10 8 5
4 Condition of Length, persistence <1m 1–3m 3–10m 10–20m >20m
discontinuities 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 5 3 1 0
Infilling (gouge) None Hard filling         Soft filling
– <5 mm >5 mm <5 mm >5 mm

Underground hard rock mining 207


Rating 6 4 2 2 0
Weathering Unweathered Slightly Moderately Highly Decomposed
weathered weathered weathered
Rating 6 5 3 1 0
5 Ground water Inflow per 10m None <10 litres/min 10–25 litres/ 25–125 litres/ >125 litres/min
tunnel length min min
ρ w /σ1 0 0–0.1 0.1–0.2 0.2–0.5 >0.5
General conditions Completely dry Damp Wet Dripping Flowing
Rating 15 10 7 4 0

ρ w = joint water pressure; σ1 = major principal stress (Continued)


208
Underground hard rock mining
Table 3.8 RMR Classification parameters (Bieniawski, 1989) (Continued)

B. RATING ADJUSTMENT FOR DISCONTINUITY ORIENTATIONS

Very favourable Favourable Fair Unfavourable Very unfavourable

Ratings Tunnels 0 –2 –5 –10 –12


Foundations 0 –2 –7 –15 –25
Slopes 0 –5 –25 –50 –60

C. ROCK MASS CLASSES DETERMINED FROM TOTAL RATINGS

Rating 100–81 80–61 60–41 40–21 <20

Class No. I II III IV V


Description VERY GOOD GOOD FAIR POOR VERY POOR

D. MEANING OF ROCK MASS CLASSES


Class No. I II III IV V
Average stand-up time 10 years for 6 months for 1 week for 10 hours for 30 minutes for
15m span 8m span 5m span 2.5m span 1m span
Cohesion of the rock mass >400 kPa 300–400 kPa 200–300 kPa 100–200 kPa <100 kPa
Friction angle of the rock mass <45° 35–45° 25–35° 15–25° <15°
Underground hard rock mining 209

Figure 3.8 T
 he relationship between the Q value and the equivalent dimension (De) (Grimstad &
Barton, 1993)

Figure 3.9 shows the relationship between RMR, roof span and stand-up time
within the limits of the RMR data set used by Bieniawski.

3.2.1.4 Mine Rock Mass Rating (MRMR)


The MRMR system was introduced by Bieniaswki and Laubscher as a further refine-
ment of the systems development (Bieniawski, 1974). The MRMR system used the RMR

Figure 3.9 RMR classification (Bieniawski, 1989)


210 Underground hard rock mining

system with some modifications and some additional adjustments. It was initially devel-
oped for caving operations but was further developed for many other mining methods.
An overview of MRMR (Laubscher, 1975) is given in Figure 3.10.
Jakubec and Laubscher (2000) caution that the following can cause errors when
applying the MRMR:

• Averaging values across different distinct rockmasses


• Mixing geological and mining-induced fractures

Figure 3.10 Overview of the MRMR components (Laubscher, 1990)


Underground hard rock mining 211

• Combining in-situ RMR with MRMR results


• Making incorrect adjustments (in alteration and weathering for example)
• Ignoring rockmass anisotropy and its orientation

The Design Rock Mass Strength (DMRS) in Figure 3.10 can be used to estimate exca-
vation stability as shown in Figure 3.11.
Table 3.9 (after Bieniawski, 1989) relates the MRMR result to a possible mining
method.
As a general cautionary point, aside from RQD, other rockmass rating systems are
prone to individual interpretation and bias, so it is critical that only very experienced
personnel are used across the whole mine to try and ensure a consistent bias, thus still
giving qualitatively accurate results.

Figure 3.11 Support requirements as a function of DRMS (Laubscher, 1990)


212 Underground hard rock mining

Table 3.9 Mining methods related by MRMR classification (Bieniawski, 1989)

MRMR Class 5 4 3 2 1
Rating
5–20 21–40 41–60 61–80 80–100

Block caving
Undercut, m 1–8 8–18 18–32 32–50 +50
Cavability Very good Good Fair Poor Very poor
Fragmentation, m 0.01–0.3 0.1–0.2 0.4–5 1.5–9 3–20
2nd lay-on blast/ 0–50 50–150 150–400 400–700 +700
drill, g/t 0–20 20–60 60–150 150–250 +250
Hang-ups as 0 15 30 45 >60
% of tonnage
Dia. of draw 6–7 8–9 10–11.5 12–13.5 15
zone, m
Drawpoint
span, m
Grizzly 5–7 7–10 9–12
Slusher 5–7 7–10 9–12
LHD, m 9 9–13 11–15 13–18
Brow support Steel and concrete Concrete Blast protection
Reinforced concrete
Drift support Lining rock Reinforced Lining Rock
techniques repair reinforcement reinforcement
Width of 1.5–2.4 2.4–3.5 2.4–4 4
point, m
Direction Towards low stress Towards high stress
of advance
Comments Fine Medium Medium to Coarse
fragmentation, fragmentation, coarse fragmentation,
poor ground, good ground, fragmentation, Large LHDs,
heavy support, fair support good drill Drill hang-ups
repairs hang-ups
Sub-level caving
Loss of holes Excessive Fair Negligible Nil Nil
Brow wear Excessive Fair Low Nil Nil
Support Heavy Medium Low Localized Nil
Dilution Very high High Medium Low Very low
Cave SI, m 1–8 8–18 18–32 32–50 +50
Comments Not Applicable Suitable Suitable Suitable,
practicable large
HW cave
area
Sub-level; open
stoping
Minimum span, m 1–5 5–20 20–30 30–80 100
Stable area, i.e., N/A 1–8 8–16 16–35 +35
SI, m
Approximate angles of pit slopes
Adjusted class 1 2 3 4 5
Slope angle 75 65 55 45 25

HW = Hangingwall; LHD = Load-haul-dump trucks; N/A = Not applicable


Underground hard rock mining 213

3.2.2 The production cycle


The drilling of rock and subsequent charging with explosives or installing support
are still the key to almost all hard rock mining operations. There are only a limited
number of exceptions and these include:

• The use of continuous miners and roadheaders in softer rock (such as coal, salt
and limestone)
• The use of tunnel boring machines (TBMs) or similar
• Certain experimental techniques (such as ultra-high-pressure water cutting,
plasma cutting or microwave-assisted breaking)

The drill and blast method is the common method for tunnelling internationally. The
method can be used in all types of rocks, and the initial cost is often lower and more
versatile than using a TBM. This tunnelling method involves the use of explosives.
Overbreak, rock fracturing around the excavation perimeter and blast vibrations are
the biggest problems that need to be controlled (after safety in general, which must
always be the first consideration). The process is not continuous and is generally lim-
ited to an excavation advance of 3–9m per day.
The typical cycle is as follows:

• Drilling blast holes and loading them with explosives


• Detonating the blast (first removing all personnel and equipment for a “re-entry
period” so the ventilation can clear the noxious fumes and dust)
• Removal of the blasted rock (mucking)
• Scaling the roof and walls/ribs to remove loose material
• Installing primary ground support
• Advancing ventilation, utilities (services) and rail (if applicable)
• Re-survey and/or geological map the area if needed and prepare the face to be
drilled by putting the hole pattern and orientation onto the face (using paint or a
laser/light system typically)

The cycle time is very important and must be planned, controlled and executed care-
fully in order to maximize the monthly excavation advance. A whole development
cycle needs to be completed each production shift or the monthly advance will be
significantly reduced.
With the development of mechanized and even automated drill rigs, the production
cycle bottleneck is the support installation. This is particularly the case in deep hard
rock excavations where seismicity is an issue, so additional and energy absorbing sup-
port is essential.
A typical eight-hour development cycle could be:

• Make safe and water down 30 mins


• Mucking (blasted rock removal) 120 mins
• Install support 120 mins
• Mark off, survey and prepare the new face drill holes 30 mins
• Drill the blast holes 75 mins
• Charge the drill holes 60 mins
214 Underground hard rock mining

Figure 3.12 Typical hard rock production cycle


Source: Railsystem.net

• Remove equipment, etc. 15 mins


• Centrally blast and re-entry period 30 mins
Total cycle time: 8 hours

The purpose of the example above is to show that if the mine uses three shifts of eight
hours, if anything is delayed, then three blasts per day becomes only two, which has
serious consequences. Figure 3.12 shows a typical hard rock production cycle.

3.2.3 Hard rock mine planning


Mine planning is usually divided into long-range planning and short-term plan-
ning, and the short-term planning shows how the long-term plan will be achieved
in detail.
Long-term planning is more strategic, broad brush and obviously involves a much
longer time period (up to life of mine). Short-term planning is tactical and becomes
more and more detailed as the period reduces from monthly to weekly and finally
to daily.
Long-term planning requires an understanding of the following in the present and
projected into the future:

• Metal or other ore prices


• Exchange rates
• Inflation rates
• Capital costs
• Working costs
• Processing efficiencies
Underground hard rock mining 215

• Local regulations, royalties and taxation


• Local political and workforce stability

Planning time periods would be (the first three typically long term):

• Life of mine
• 5-year plans
• Annual plans
• Quarterly plans
• Monthly plans
• Daily activity plans

The planning process is a key part of a mine business plan that in varying detail
includes the product schedule to meet the plan. It ensures that the profitability of
the orebody is maximized over the life of mine, taking into account the changing
parameters (especially the ore commodity price). This is a very complex and changing
process that is ideally suited to the use of an appropriate computer planning software
package. It typically involves considering the following:

• Capital requirements
• Operating costs
• Revenue
• The block orebody model (including cut-off grade)
• The mine design
• Optimum mine production rate, sequencing and scheduling

Planning is and must be a dynamic process that is continuously revisited and updated.
The mine planning objectives are shown in Figure 3.13 (modified from Frimpong,
2011).
The Block Model is the key to mine planning and is developed from the Geology
Model. An orebody is divided into numerous layers of blocks all usually having the

Figure 3.13 Mine planning objectives (modified from Frimpong, 2011)


216 Underground hard rock mining

Figure 3.14 A block model below an existing openpit


Source: GEOVIA, Dassault Systemes, 2020

same volume and dimensions, in all three axes (x, y and z). Selecting an optimum
block size is not a trivial decision, and specialized planning-related software models
can assist with this critical decision. Each block is assigned unique properties, the
most important of which is the block grade. Other parameters could include the den-
sity, rock type(s) and rockmass properties.
Figure 3.14 shows a block model below an existing openpit.
Calculating a reliable grade for each block from generally just core holes is a real
challenge. Krige (1951) developed a geostatistics method (referred to as Kriging) to do
this, using a Gaussian regression process to interpolate between individual borehole
grade results.
According to the Environmental Systems Research Institute (2016), Kriging assumes
that the distance or direction between sample points reflects a spatial correlation that
can be used to explain variation in the surface. The Kriging tool fits a mathematical
function to a specified number of points, or all points within a specified radius, to
determine the output value for each location. Kriging is a multistep process; it includes
exploratory statistical analysis of the data, variogram modelling, creating the surface
and (optionally) exploring a variance surface. Kriging is most appropriate when you
know there is a spatially correlated distance or directional bias in the data. It is often
used in soil science and geology.
Stochastic modelling techniques have proved essential in the mine planning process.
According to Wikipedia, “stochastic” refers to the property of being well described
by a random probability distribution. Although stochasticity and randomness are dis-
tinct in that the former refers to a modelling approach and the latter refers to phenom-
ena itself, these two terms are often viewed as being synonyms. Stochastic Kriging
has been developed and is a useful tool in creating block models that are essential for
effective mine planning (Ankenman et al., 2010).
Underground hard rock mining 217

3.2.4 Mine planning challenges and uncertainties


As more information becomes available due to mining, and costs and revenues change
(from metal price variations for example), planning must be updated regularly.
All mining operations require accurate, reliable and timely data. Planning requires
considerable and diverse data that are all prone to errors and uncertainty. There are
many parameters needed for accurate planning that may be unreliable because of the
variability within the rockmass and the interpretation, including:

• The geological model (especially the dimensions of the orebody and the grade)
• The rock mechanics model (which defines the mine design and support)
• The block model (especially the grade distribution)

Dominy et al. (2002) list the five main geological reasons for errors in the block
model:

• Poor and unreliable sampling and assay data


• A lack of detailed information on the orebody geology
• Poor interpretation of the grade and its distribution within the orebody
• Poor understanding, usage and interpretation of relevant software
• Failure to recognize the volume variance effect (usually resulting from too much
dependence on small samples rather than large ore block grades)

Dominy et al. (2002) further state that the grades in block models (resource and reserve
estimates) are in practice downgraded based on Due Diligence Studies and Audits.
This they state is due to one or several of the following issues:

• Core drill hole relative orientation (direction) compared to the ore zone mineral-
ization dominant orientation
• Inadequate primary sampling
• Assay accuracy and repeatability
• Poor correlation between analyses of duplicate field splits. A field split is thor-
oughly homogenized and then divided into generally two samples.
• Poor, inconsistent or variable core sample recovery
• Biased sampling techniques
• Inappropriate and/or mixed drilling methods (e.g., wet reverse circulation
drilling)
• Poor correlation between analyses from twinned holes (a hole drilled near an
existing one to confirm the results)
• Down-hole contamination
• Lack of orientation surveys particularly in long holes
• Poorly understood geological and grade continuity
• Inappropriate geological interpretation and modelling techniques
• Inappropriate resource estimation techniques
• Poor dilution and loss estimation
• Incorrect determination of the ore and waste bulk density
• Impractical mine planning assumptions
• Unforeseen metallurgical recovery issues
218 Underground hard rock mining

3.3 Mine access and development

3.3.1 Introduction to shaft and decline access


The main underground development openings are designed so that ore bodies are
safely, easily and efficiently accessible for men and materials. The development
would also need to ensure that the ore can be moved to the surface for process-
ing effectively. Access can be typically categorized into three, depending on their
importance:

• Primary or main openings (e.g., shaft, adit or slope/decline/ramp)


• Secondary, level or orebody openings (e.g., haulage, drift or entry and ore-pass)
• Tertiary or lateral or panel openings (e.g., internal ramp or crosscut)

The construction of underground openings is a specialized operation, time consum-


ing and costly. Mine development has therefore become increasingly mechanized and
efficient in order to reduce costs and time to develop and start mining a deposit.
The most critical decision is the mine access method from surface and the location.
A number of other decisions related to the primary development openings for a mine
must also be made early in the mine planning stage and include the type, number,
shape and size of the main openings. Factors that influence this decision include:

• The depth, shape and size of the deposit


• The surface topography
• The geological and hydrological conditions of the orebody and surrounding rock
• The mining method and the production rate
• The climate
• The local infrastructure available
• Relevant regulations
• Sustainability and environmental considerations

The size of the main access development depends mainly on:

• Ventilation requirements
• The tonnes that must be hoisted per day
• The size of mobile equipment to be used
• The underground personnel required
• The provision of services (electricity, water, compressed air, etc.)

Primary underground development openings can also be used for exploration pur-
poses. Those openings, if driven in advance of full-scale mining, can provide valuable
additional and detailed exploration information and afford suitable sites for future
exploration drilling and sampling. Obviously, excavations driven for exploration pur-
poses can be utilized to develop the deposit; some shafts, adits, declines and drifts
would later serve to accelerate the opening up of a new ore deposit.
Underground mines are used for accessing and exploiting ore bodies that are not
close enough to surface or that have already been mined from surface and are still
viable at depth, but uneconomic for continued surface mining.
Underground hard rock mining 219

Figure 3.15 General layout of an underground mine (Hamrin, 1982, 2001)

The infrastructure of underground mines is obviously more complex than that of


surface mines (open-pits or strip mines). A typical layout of an underground mine is
shown in Figure 3.15.
A main feature of an underground mine is often the shaft (or shafts), which is a ver-
tical or subvertical access to the underground workings, or depending on the depth
and extent of the deposit, it can be an inclined tunnel, called a decline or ramp. Both
types of access developments, shaft and decline, can be present in the same mine,
as shown in Figure 3.15. In mountainous terrain, the underground orebody can be
accessed from the slope of a hill (or a mountain) using horizontal or gently inclined
tunnels, called adits. The surface entrance to a decline or an adit is called a portal.
Generally, horizontal excavations are developed slightly up-dip to facilitate gravity
drainage away from the advancing excavation.
Underground mines usually have at least two independent access systems because
of safety considerations for the workforce underground if one becomes inaccessible
for some reason, and to facilitate ventilation. In many countries, this is required by
law. This can be achieved by using various combinations such as a ramp for equip-
ment, personnel and intake air and a shaft for transporting ore out of the mine and for
upcast (waste) ventilation.
There are generally three methods of accessing an underground mine: shaft, adit
and decline or ramp. The shaft often remains the mine’s most critical infrastructure,
and downward development is frequently via ramps to allow access for the mobile
equipment. A decline ramp from surface can facilitate easy machine movement and
220 Underground hard rock mining

Figure 3.16 Decision tree for main access transportation determination (De La Vergne, 2003)

the transportation of people and materials. It can also be used for ore transportation
by truck or conveyor, eliminating the need for hoisting shafts, but that is an economic
decision as shafts are typically more effective and cheaper for moving especially large
tonnages to surface especially from deeper mines.
The selection of whether a shaft or ramp is more effective as the primary access
depends mainly on the orebody depth and planned production rate. De La Vergne
(2003) proposed the decision tree as shown in Figure 3.16.

3.3.2 Shafts, raises, adits and ore (rock) passes

3.3.2.1 Shafts
A shaft is a vertical excavation (but could be inclined too, although this option is
seldom used now as it is less efficient) in which elevators are used to transport people,
equipment and ore (and sometimes waste) in and out of the mine. Shafts are always the
option used where the deposit is located at depth.
Shafts are generally used for the following functions:

• To access an orebody
• To transport men and materials to and from underground workings
• For hoisting ore and waste from underground
• To serve as intake and/or return airways for the mine (ventilation)
• To provide a second egress as required by mining law
• Access to nuclear waste storage
• Hydropower generation
• Access underground civil structures such as basements and underground rail sta-
tions or road tunnels
Underground hard rock mining 221

Most shafts are divided into a number of compartments, each with a different use,
by brattice walls or steel works fitted with conveyance guides. For example, one major
compartment could be used for moving people and equipment, a second with two
skips for taking rock to the surface and other compartments for service infrastructure
(typically pipes and cables) and space for ventilation. The main factors to establish
the shaft size are the monthly tonnage requirement and the ventilation needed. At
the mine planning stage, it is useful to include some excess rock handling and venti-
lation capacity in case future mining production increases. The orebody size, grade
and mining method will determine the rate of mining, and thus the tonnage (ore and
waste) to be hoisted, the size of the workforce and the material to be moved efficiently
in any given shift.
Once a shaft is excavated to its final depth, typically some distance below the eco-
nomic depth of the orebody for ore handling and water handling, the shaft is equipped
with all the necessary steel work, etc. to guide the shaft conveyances and hold the
pipes and cables for the services needed underground. Figure 3.17 shows a cross-sec-
tion through a typical shaft.
Shafts are often essential for underground mines, and their location is determined
based on detailed surface topography and infrastructure, the orebody, geology, rock
mechanics and environmental assessments. The location must be changed where
adverse geotechnical conditions are identified in the originally planned site. Proximity
to the orebody, strata conditions and water-bearing structures are the major parame-
ters that govern the ultimate location of shafts. The decision to locate the shaft is crit-
ical because the process to develop a shaft is very expensive and relatively slow. The
vertical shaft must be well located with respect to the ore deposit and to be able to han-
dle production needs. The correct diameter and configuration of the shaft will provide
optimum operational efficiencies. The shaft can be rectangular, circular or elliptical
in profile, although almost all hard-rock underground mines have circular section
shafts because this shape generates a better geometry for airflow and is naturally more

Figure 3.17 Sections of a typical shaft layouts (Kempson et al., 2015)


222 Underground hard rock mining

stable from a stress concentration point of view. At great depth, however, an elliptical
profile can be used if one of the principal horizontal stresses is significantly larger than
the other. Elliptical shafts were designed as an alternative to large circular shafts by
simply adding half-moons along the main axis. This had the effect of reducing the cir-
cular excavation and therefore the cost of sinking the shaft and could be useful under
certain high horizontal stress conditions (where σ2 >> σ3).
Most shafts constructed in the early 1900s were of a rectangular cross-section
mainly because they were easier to equip with square or rectangular conveyances using
timber beams and were not so deep that stresses caused instability issues. Blasting a
square or rectangular cross-section was, however, problematic and this slowed down
the rate of sinking.
Circular shuttering for concrete lining is easier to move when doing concurrent lin-
ing resulting in faster work progress during sinking operations. This is an important
aspect when it comes to the cash flow for any mining project.
The shape and size of equipment to be taken down a shaft are also considered in the
calculation of the final shaft dimensions. This process is equally applicable to ramps
(declines), the main difference being that the ramp is equipped to rather handle track-
less equipment and possibly be fitted with conveyors instead of the skips and cages.
Inclined shafts can also use monorails.
Determining the rate of mining can be summarized as follows:

• Identify possible mining methods


• Define standard mining blocks (stope or panel size) per method
• Calculate steady state conditions per level
• Define steady state inputs/outputs requirements per level
• Determine minimum access dimensions to cater for equipment and ventilation
• Calculate development requirements to get to steady state
• Simulate full level production from start of block to orebody extremity
• Determine the maximum number of levels that will operate simultaneously
• Estimate shaft size required to cater for the sum of the requirements of the maxi-
mum number of working levels
• Do economic analyses using, for example, the net present value (NPV) and inter-
nal rate of return (IRR)
• Decide on optimum mining layout and shaft configuration
• Ensure an adequate and achievable rate of investment is achieved; if not, repeat
the process

3.3.2.2 Raises
Raises are steeply inclined openings linking the mine sublevels at several vertical eleva-
tions and are often developed within the orebody. They are normally placed near or in
the stopes employing specialized cyclic or continuous operations. Specific applications
of bored raises are transfer of material, ventilation, personnel access and ore production.
Inclination varies typically from a minimum of 55° to the vertical, which is the lowest
angle of repose of blasted rock. They have variable cross-sections from typically 2 to 30m2.
Since manual excavation of raises is a potentially more hazardous activity, raise
boring is frequently utilized for developing ventilation raises, ore passes and rock fill
passes. It provides safer and more efficient mechanized excavation of circular raises
Underground hard rock mining 223

up to 6-m diameter because this method eliminates the need for explosives, and keeps
personnel remote from the actual rock excavation.

RAISE BORING

Raise boring is the procedure of mechanically boring a vertical or inclined shaft between
two or more levels (one of which could be the surface). Access must be available at the
start and end of the excavation. In conventional raise boring, a downward pilot hole
is drilled to the target level by the raise boring machine, where the bit is removed and
replaced by a reaming head (Figure 3.18). The development of directional drilling has
made raise boring long shafts even more feasible for longer excavations as the pilot hole
can be kept correctly aligned and positioned. After the pilot hole is complete, a reamer
head is fitted and the machine then reams back the hole to final diameter, rotating and
pulling the reaming head upward towards the start of the pilot hole (where the raise
boring machine is securely located). The cuttings fall to the lower level and are removed
by any convenient method using remote control for safety considerations. The capital
cost of a raise boring machine is relatively high, but the return on investment is good
particularly for longer length excavations. Advantages of raise boring are that miners
are not required to enter the excavation while it is underway, the excavation process is
continuous, no explosives are used, a smooth and more stable profile is obtained and
manpower requirements are reduced. Safety is, however, the most important advantage.

Figure 3.18 Raise boring process (Atlas Copco, 1997)


224 Underground hard rock mining

ALIMAK RAISING (www.alimak.com/industry/mining/)

Alimak raising is another excavation alternative to raise boring that is still safer than
conventional raise development. For ore-passes (rock-passes, box-holes), short raises
and short shafts, a semi-mechanized system can be used especially if only bottom
access is available. The excavation must be vertical or steep dipping to use this system,
which was developed in 1957 by Alimak. It was initially developed (and still used) for
high rise building construction.
Typical mining applications include:

• Ventilation raises
• Stope slot raises (for longhole open stope first production, for example)
• Ore and waste pass systems
• Raise bore rehabilitation
• Ventilation shafts
• Manway raises
• Emergency ladder way installations

Alimak Raise Climbers are available with air, electric or diesel/hydraulic drive units.
One of the longest shaft driven in one step was 1,050m long in Norway. The platform
can be any shape and size. The largest work platform supplied so far measured more
than 30m2.
The raise climber system is basically a movable working platform that runs along
a monorail beam that increases productivity and improves safety when compared
with other conventional blasting systems. As the excavation is developed, the mon-
orail track is lengthened by adding sections to it and bolting them to the excavation
sidewall.
Using the system involves five production steps (as shown in Figure 3.19):

• Scale (make safe from the protection of the personnel cage fitted with a protective
canopy)
• Drill (conventionally from the cage)
• Charge the holes (then remove the cage from the excavation using the monorail,
which has a drive-toothed section)
• Blast (remotely)
• Ventilate and remove the broken ore from the bottom of the excavation

Figure 3.19 The Alimak raise development production steps


Source: Vikay Mining Equipment
Underground hard rock mining 225

Step 1 Drilling
• Drilling is undertaken from the drill deck on top of the raise climber, which
is sized to suit the size, shape and angle of the raise.

Step 2 Charging-up
• When drilling is complete, the face is charged with explosives.

Step 3 Blasting
• The Alimak climber is then lowered to the bottom of the raise and into a
station for protection before the blast is triggered from a safe location.
• The rail is left in place and protected (only the top is really exposed to the
blast directly).
Step 4 Ventilation

Step 5 Scaling (removing lose and unstable rocks)


• The Alimak system provides for efficient post-blast ventilation and a powerful
air/water blast effectively dislodging loose rock from the freshly blasted face,
making it ready for re-entry.

3.3.2.3 Adits
An adit is a near-horizontal excavation that is used in mountainous areas or from high-
walls where the orebody is located near or above the access (valley) floor (Figure 3.20).
They are typically developed slightly up-dip at about 4° so that any water will drain
out of the adit under gravity. Developing an adit is the same process as developing
any horizontal tunnel and is an option only where the topographic relief is con-
siderable. In this access opening, the ore and waste can be taken out of the mine
at minimal operating cost. The traditional method of developing adits is to drill
and blast the face, load the material into a haulage device (after the area has been cleared
of dust and fumes) and then provide support and ventilation to the newly advanced
face. Thus, drilling and blasting are the standard excavation method for adits and most

Figure 3.20 Footwall adits development, generalized (Zhang & Wang, 2016)
1. Adits 2. Haulage drift 3. Ore pass
226 Underground hard rock mining

Figure 3.21 Decline development (Zhang & Wang, 2016)


1. Decline 2. Crosscut 3. Sub-level drift 4. Ore-body

hard-rock horizontal development. TBMs can be used depending typically on the length
and the economics. The other main exception to the use of drilling and blasting are under-
ground mines in relatively soft rock such as coal, salts and limestone where the rock can be
removed without the need for blasting using continuous miners or road headers.
A decline or ramp is an access tunnel usually developed at a low slope angle from the
horizontal (<20° dip) (Figure 3.21). The design and support of declines are considered
the main challenges in underground mine development as they are generally needed for
the life of the mine. They can be mainly straight, spiralled or a combination of both.
Ramp access is the common selection in shallow ore bodies. A ramp from surface can
facilitate machine movements and transport of people and materials. It can also be used
for ore transportation by truck or conveyor, eliminating the need for hoisting shafts.
Ramps are sized to include space for equipment and must incorporate a safe margin for
clearance, walkways, ventilation and other facilities and services. Cross-sections typi-
cally vary from 2.2m × 2.5m in mines with a low degree of mechanization and low ton-
nage to 5.5m × 6.0m where large equipment is used and high ore tonnages are needed.
In many mines, the decline is used to transport ore to the surface using a conveyor belt,
being associated with grade limits. For example, if utilized for conveyor belt haulage
only, the maximum grade of the decline could be between 10° and 15° depending on the
material to be conveyed. If trucks are used, the typical decline is between 8° and 10°.

3.3.3 Main access development


Near the shaft or decline, service excavations are typically developed for maintenance,
offices, pump chambers, etc. After accessing the orebody, underground workings are
continued on successive near-horizontal planes, referred to as levels, each level being a
system of related underground workings located approximately on the same horizon-
tal plane. The underground workings include production stopes and mine development
infrastructure for transportation of men and materials and the broken ore (and some-
times waste) to the surface for processing (or disposal). Between the main development
levels, a mine can have sublevels, which may be needed for more effective drill and blast
control in the stopes. Levels and sublevels are mainly connected by inclined underground
openings, also called ramps, and also by the vertical opening, including raises and win-
zes. The main elements of the underground infrastructure are shown in Figure 3.22.
A drive or drift is a horizontal or nearly horizontal underground opening devel-
oped on the underground levels along the strike of the orebody. The drives are
subdivided into hangingwall (located at the upper ore-waste contact, “hanging”
Underground hard rock mining 227

Figure 3.22 Basic infrastructure and terminology for an underground mine (Hamrin, 2001)

above the orebody) and footwall (located at the lower ore-waste contact, at the
“foot” of the orebody) drives. A footwall drive is also commonly called an ore
drive. In many hard rock mines, they are developed in pairs, one called the haul-
age that is used for men, materials and intake (fresh) ventilation and the other the
return air way (RAW).
A crosscut is a horizontal underground tunnel developed to intersect the orebody.
The crosscuts are usually developed to connect the drives with the area in which stop-
ing (ore production) occurs.
Depending on the nature of the orebody, a raise is an underground opening driven
upward and a winze is driven downwards typically in the orebody.
A stope is an underground excavation where the actual ore is produced. Development
of stopes in massive mining methods often starts from blasting a slot, which is a steeply
dipping excavation at one boundary of the planned orebody stope. Mining then con-
tinues by blasting rings or slices of the orebody into the slot.
228 Underground hard rock mining

A pillar is a block of ore or barren rock left intact in the mined-out stope, between
two stopes or between two drives to act as a regional support. It is required to provide
structural integrity to the stoping process and prevent the stope walls from unplanned
collapse. Pillars may be extracted after stopes are mined out, but some pillars may be
left in place permanently to provide regional support.
A drawpoint is a place from which the ore is extracted from the stope and loaded
onto trucks or conveyors for further transportation to surface.
An ore pass is a steeply dipping (+55°) underground opening for moving mainly ore
from one level to a lower level under gravity and finally to the shaft bottom from where
it is hoisted to the surface. The ore is loaded through the chutes, which are the loading
arrangements that utilize gravity. An important element of the ore loading and trans-
portation system is a coarse steel grating, called a grizzly, for screening out oversized
rock fragments that could block the ore passes.
In an underground mine, a significant amount of infrastructure must be built before
mining begins, which requires a very significant capital investment (see Figure 3.22). The
development of a large underground mine can take as many as 5–10 years. Costs during this
time will therefore be high and can comprise 30–40% of the pre-production capital require-
ments before mining can start. In these cases, a dual feasibility study must be performed
comparing the surface option to the best underground mining option. Generally, capital
costs increase and operating costs decrease with increasing production tonnage targets.
An underground mine has a unique layout that ensures the cost-effective extraction
of ore and the safety and movement of people and equipment. Therefore, each mining
method requires different underground infrastructure such as access drifts to sub-
levels, drifts for longhole drilling, loading drawpoints and ore passes. Together, they
form an intricate network of interdependent excavations such as drifts, ramps, shafts
and raises. The mine requires three different physical installations:

• The surface process plant and infrastructure


• The shaft and/or ramp infrastructure
• The underground infrastructure

The first consists of a variety of facilities to provide the mine with necessary services
such as: access roads and parking, transportation facilities, power and water sup-
ply, service and maintenance buildings, mineral processing plant, bulk storage and
waste disposal facilities for air, water and solids. The shaft plant includes the facilities
installed for material handling of ore and associated waste and the means of transport
of miners and material. It generally incorporates systems for rock handling, ventila-
tion, water handling, services and power supply, and communications.
The underground plant covers various installations to make the system work effi-
ciently and safely, including storage bins (silos), loading pockets, rock handling systems,
power distribution equipment, underground maintenance facilities and numerous other
installations that provide auxiliary services to the underground operations.
Mine ventilation is one of the most important facilities of underground mining. Air
quality in mine workings is an area of particular concern to the underground develop-
ment. It must be maintained at an acceptable health standard. A continual and ade-
quate supply of fresh air must be made available to working areas. Ventilation shafts
are shown in Figure 3.23. Underground mines use networks of fans, gates and surface
openings to move fresh air into the mine and remove exhaust air. High-pressure fans
Underground hard rock mining 229

Figure 3.23 General layout of an underground mine development (Zhang & Wang, 2016)
1. Shaft 1; Skip shaft 2. Shaft cage for moving men, materials and equipment 3. Crosscut
4. Shaft station development 5. Rock or ore pass 6. Primary crusher 7. Ore bin
8. Ventilation shaft 9. Haulage level 10. Stope (production area)

on surface extract exhaust air through the upcast shafts and ventilation doors control
the underground airflow, passing fresh air through active work areas. As most of the
infrastructure is located on the footwall side of the orebody, the fresh air is normally
channeled via the footwall toward the hangingwall, from where the exhaust air is routed
to the surface. It is particularly important to clear the air after an underground blast,
because harmful gases such as carbon monoxide or oxides of nitrogen can build up. A
good ventilation system will rapidly clear the air from a blast of dust and noxious gases.

3.3.4 Main access protection


The natural strength of the rocks is seldom sufficient for the safe and unsupported
excavation of the rockmass in the underground mines. In order to prevent a rockfall,
the rock around the excavation must be reinforced (laterally confined) using pillars
and installed support.
The installed support is typically rockbolts and cablebolts often with skin support
of mesh (screen) or sprayed concrete (shotcrete).
The main access, ramps and especially shafts must be protected for the life of the mine.
The protection can require more than just installed support and may need pillars too. Even
where shafts are completely stable, a cosmetic (concrete) lining may be installed to reduce
friction losses in ventilation or to hang shaft hardware for hoisting, support of utilities, etc.
Shafts with hoisting need special protection from mining as movement, especially
shear or tilt, cannot be easily accommodated whilst maintaining high and efficient
hoisting speeds.
230 Underground hard rock mining

3.3.4.1 Shaft protection in a massive orebody


If the orebody is massive, generally the shaft is sited in the footwall some distance
from the deposit so the mining-induced stresses don’t cause future stability issues. The
exception to this is where the footwall strata is potentially unstable and the hanging-
wall is much more competent.

3.3.4.2 Shaft protection in a tabular orebody


Tabular deposits therefore need a unique approach, and there are only a few methods
to protect a shaft from subsequent damaging effects induced by mining. These are:

• Site the shaft away from the orebody (more costly and time-consuming access to the
deposit and long transport distances over the mine life for men, materials and rock)
• Leave a shaft pillar (can become uneconomic especially at great depth)
• Leave satellite pillars
• Mine out the orebody concurrent with the shaft sinking (early pillar removal)

For deep shaft pillar design, the main criteria to consider are the vertical strain, vertical
tilt and the vertical stress.
The following were postulated by Wilson and More O’Ferrall (1970):

• Shaft damage begins when the field stress exceeds a quarter of the UCS of the
surrounding rock.
• When the field stress increases to about half the UCS, damage becomes extensive
and remedial actions very costly.
• When the field stress becomes two-thirds of the UCS, it is impractical to keep
excavations open.

Salamon (1974) concluded that the maximum tolerable vertical strain was 1 millistrain
and Wagner (1980) reported a maximum tilt of 1 millistrain.

3.4 Mine method selection

3.4.1 Mine design considerations and methodology


There are many factors that influence the selection of a mining method, and these have
not changed significantly over time (Warner, 1934). The factors can be categorized as
follows:
Internal factors such as:

• Orebody
• Ore
• Walls and capping

External factors:

• The mining location


• Climate
Underground hard rock mining 231

Local economic factors:

• Labour
• Services
• Material/equipment availability
• Capital availability
• Local infrastructure

Global economic factors:

• Demand and price fluctuation of the commodity


• The effect of the new mine on supply/demand

Orebody factors such as:

• Shape and size


• Horizontal area of commercial ore zone
• Width-to-length relation
• Vertical extent
• Boundary regularity
• Attitude
• Dip and/or pitch
• Relation to surface and public/private land
• Relation to other economic orebodies or mines
• Continuity
• Presence of economic off-shoots, etc.
• Presence of waste intrusions
• Presence of major geological features traversing the orebody
• Presence of abnormally wide or narrow sections

Specific ore factors such as:

• Value per unit of the ore


• Strength and macro-structure
• Value distribution
• Uniform
• Sorting requirement if applicable (waste, high grade or different metallurgical need)
• Need for selective mining or more than one method
• Ease of concentration
• Mineralogy and micro-structure
• Packing (in rock passes or stopes)
• Breaking
• Explosive required
• Fragmentation
• Behaviour if undercut or crushed
• Weathering
• Swell
232 Underground hard rock mining

• Oxidation
• Health hazard creation (e.g., gases, dust, etc.)
• Explosion or fire hazard creation (e.g., methane, hydrogen sulphide, etc.)

Wall and capping factors such as:

• Strength (support and subsidence)


• Reaction to stress
• Failure nature (coarse/slabs/blocks or fine, free running or sticky)
• Regularity and dip (see orebody)

Mining location factors such as:

• Topography as it affects:
• Transportation
• Infrastructure location (shafts, workshop, plant, etc.)
• Ore storage
• Materials storage
• Waste disposal
• Water drainage
• Underground water system
• Source, quantity, quality and temperature
• Water table level
• Water-bearing strata
• Impervious strata
• Seasonal variations
• Presence of major structural features
• Presence of major surface water sources (e.g., rivers, lakes, wetlands, etc.)
• Temperature gradient/variation on surface and underground

Local economic factors such as:

• Labour
• Availability
• Skills
• Whether organized
• Political stability
• Services (mainly power)
• Supply, reliability and cost or electricity
• Supply, reliability and cost of fuel, etc.
• Capital availability
• Scale of operation
• Time from start to production
Underground hard rock mining 233

• Local infrastructure
• Distance to towns and resources (schools, hospitals, etc.)
• Size of nearby towns
• Local quality of life
A relatively new factor is the effects of mining on the local community as this can
influence the mining method selection.

3.4.2 Rockmass classification


A suitable classification method is vital especially in the early stages of justifying the
mining of an orebody. This has already been covered in Section 3.2. This helps deter-
mine one or more candidate mining methods that can be further investigated and
refined.

3.4.3 Basic mining method options


The final mining method selection depends on many factors; therefore, and because of
varying orebody conditions, mines could utilize more than one method or modify a tra-
ditional mining method as the need arises. The bottom line is that the mining method
selected must be safe, minimize environmental and community impact and be cost
effective.
Final considerations include:

• Determination of possible mining methods


• Estimate production tonnage possible (also requires stope size and configuration)
• Determine support requirements
• Estimate mining efficiency (% extraction of total resource, any dilution, etc.)
• Estimate total working costs
• Establish economic cut-off grade for each
• Optimize process plant and recoveries
• Determine mine access requirements
• Conduct an EIA (focussing on the local community’s needs)
• Select optimum mining method(s).

Table 3.10 shows an example of a mining method selection chart, based on various
parameters.
Brown (2003) categorized the underground mining methods as follows (Table 3.11).

3.5 Unsupported mining methods


Unsupported methods consist of:

• Longwall mining (without backfill)


• Sub-level caving
• Block caving
234 Underground hard rock mining
Table 3.10 Geological and mechanical criteria in large-scale mining methods (simplified after Brown, 2003)

Orebody characteristics Orebody configuration

Ore strength Waste strength Beds Veins/Ore Massive Ore dip

Mining method Weak Mod Strong Weak Mod Strong Thick Thin Narrow Wide Flat Mod High

Room & pillari X X X X X X X X


Longwall X X X X X X X X
Sublevel stopingii X X X X X X X
Shrinkage X X X X X X X X
Cut & fill X X X X X X X X X
Undercut & fill X X X X X X X X
Sublevel caving X X X X X X X X X
Block cavingiii X X X X X X X X

Notes:
i. A reasonably uniform thickness
ii. Fairly regular orebody contacts
iii. Strong fractured rock also can be caved
Underground hard rock mining 235

Table 3.11 Mining method classification (Brown, 2003)

Naturally supported methods Artificially supported mining methods Unsupported mining methods

Room and pillar Drift and fill Longwall mining


Sub-level stoping Cut and fill Sub-level caving
Longhole open stoping Undercut and fill Block caving
Shrinkage mining
VCR

The following notes refer to Table 3.11.


Note 1:

Moving from naturally supported to artificially supported and finally unsupported mining methods creates a significant
increase in country rock displacements.
Note 2:

Moving from unsupported to artificially supported and finally naturally supported mining methods creates a significant
increase in strain energy stored in the near field rockmass.

3.5.1 Longwall hard rock tabular mining (based largely on Spearing, 1995)

3.5.1.1 Introduction
Modified longwall methods can be used to mine narrow, flat-dipping, metalliferous
orebodies of large areal extent.
The method is used in the following typical applications:

• Limited to deep hard rock relatively flat (dip <30°) tabular orebodies
• In geologically relatively undisturbed strata
• Where the reef width is less than 5m and ideally less than 2m

The more conventional method would be room and pillar, but this method is used
when the value of the ore is high (e.g., typically gold or platinum) and leaving pil-
lars is therefore not cost effective. It is not the same caving principle as for longwall
coal mining and only refers to the mining stopes being aligned to help control the
stresses. The near-field rock is usually strong, and mining often takes place at con-
siderable depth where in-situ stresses are high. Seismicity or rockburst phenomena
can be associated with slip on discontinuities or with the stress-induced fracturing
that is commonly observed to occur around longwall hard rock (brittle) faces in
high-stress settings.

3.5.1.2 A typical hard rock longwall mine


An alternative to longwall hard rock gold or platinum mining is a scattered min-
ing layout. At depth, a scattered mining layout creates high stress remnants as the
stope (production) faces are mined towards each other creating a highly stressed
abutment that can be prone to seismicity. A typical scattered mining layout is
shown in Figure 3.24.
236 Underground hard rock mining

Figure 3.24 A typical scattered mining layout

The advantages of longwall mining over scattered mining are:

• Fewer highly stressed remnants left


• Off-reef development follows in the footwall in de-stressed conditions (in the
stress shadow created by the mining faces above)
• More concentrated production making some mechanization possible
• Only a single “starter” raise on reef between the levels is needed

The disadvantages are:

• Inflexibility (no selective mining possible)


• Less advance geological knowledge
• Difficult to handle significant reef displacements caused by faulting

A typical longwall configuration at depth, with mining being done in only one direc-
tion, so it helps reduce the stresses as shown in the plan in Figure 3.25.
The access to the longwall stope is typically as follows:

• Main access tends to be developed deep (20–40m) in the footwall (below the
reef) as the hangingwall is unstable due mainly to high induced stresses above
a stope.
Underground hard rock mining 237

Figure 3.25 Longwall configuration at depth

• Main haulages are developed usually parallel to strike and then crosscuts at right
angles stopping near the reef.
• Raises are developed to the reef plane and then on the reef to connect with the
next level.

There are two main deep tabular mining configurations: Overhand and Underhand,
as shown in Figure 3.26.
The individual stope panels are typically 30–40m long, and mainly because of the
very hard and brittle rock, the low production per stope panel and the relatively low
mining height, mechanization is very challenging. A typical panel layout is shown in
Figure 3.27 and an actual panel is shown in Figure 3.28.

3.5.1.3 The main hazards associated with hard rock longwall mining
A rockburst is defined as the uncontrolled and sudden disruption of rock associated
with a violent release of energy that is additional to that derived from gravity falling
rock fragments. The reinforcement system installed that is used to limit the effects of
rockbursts must therefore absorb energy.
238 Underground hard rock mining

Figure 3.26 The two main hard rock tabular mining layouts

The theory of elasticity and energy release rate (ERR) have been used with some
success to develop an understanding of the causes of rockbursts in longwall mining in
hard rock and to develop mining strategies that limit the incidence and effects of rock-
bursts. The stresses and displacements induced by the creation of a new excavation, or
by the extension of a longwall stope, may be calculated most conveniently using one
of the forms of the boundary element method (computer software).

ENERGY RELEASE RATE

ERR is a calculated (not a physical) parameter that considers the stresses act-
ing on the stope (face) abutment and the volumetric closure in the stope behind
the advancing face. Most of the ERR is released in a non-violent manner in the
form of fracture growth, rock crushing, fracture sliding (shear) and temperature
increase.
Cook and Salamon (1966) defined ERR as:

ERR = ∆USA /∆a in MJ/m 2


Underground hard rock mining 239

Figure 3.27 Typical stope panel layout (Spearing, 1995)

Figure 3.28 A tabular hard rock platinum stope (showing the gully and narrow stope panel)
Reproduced with kind permission from Impala Platinum Holdings Ltd.
240 Underground hard rock mining

Where:
ΔUSA is the available strain energy which relates to the stresses
Δa is the area being mined
A mining energy balance is shown in Figure 3.29.
The relationship between increased ERR and increased seismicity is only reliable,
but ERR is still a useful qualitative tool in the optimization of mining layouts as it
relates directly to the magnitude of the stresses (Cook and Salamon, 1966).
Where stabilizing pillars are used, the average pillar stress (APS) is often consid-
ered, and as a guideline, it should be limited as follows:
APS ≤ 2.5 * UCS

Where:
UCS is the unconfined compressive strength of a core sample tested in the laboratory.

Figure 3.29 Energy balance for a mining cycle (Denkhaus, 1963)


Underground hard rock mining 241

Obviously other considerations are important, such as:

• The strength of the immediate hangingwall (roof) and footwall (floor).


• The effective height of any pillar rather than the actual height needs to be consid-
ered due to the very fractured nature of the rock around the pillar and stope. This
can be as much as six times the actual mined height.
• Excess shear stress (ESS) and seismic potential

The ESS is the theoretical shear stress in excess of that required to initiate slip on a
fault or fracture.
It is based on the Coulomb-Navier failure criterion and is mainly applied to geo-
logical faulting in the general three-dimensional case of rocks containing arbitrarily
oriented strength anisotropies and subject to non-Andersonian stress systems (i.e.,
with none of the principal stresses acting in a vertical direction). The relationship is
shown in Figure 3.30.
The concept of ESS is shown diagrammatically in Figure 3.31.
ESS is helpful when designing bracket pillars against potentially seismically active
geological features such as faults.
Where a stiffer rock intrusion like a dyke exists, another basic strength to stress
criterion can be useful:
( σ1 / σc ) ≤ 1

Where:

σ1 is the maximum principal stress


σc is the UCS of the stiff intrusion

There are three basic modes of rock mass instability leading to rockbursts:
• Fault-slip events resemble natural earthquakes and usually occur on a mine panel
or mine scale.

Figure 3.30 The Coulomb-Navier failure criterion


Note: φ is the angle of internal friction and μ is Tan φ.
242 Underground hard rock mining

Figure 3.31 The concept of ESS (COMRO, 1988)

• Stress-induced fracturing or crushing of pillars, or fracturing at or near the min-


ing face, can lead to local instabilities, sometimes called strain bursts, which
occur on a stope or excavation scale.
• Stiffer rock such as dykes can store more energy than the surrounding rockmass,
and this can lead to rockbursts.

There is clearly a potential for much greater release of energy in a fault-slip rockburst
than in a pillar burst due to the larger volume of rock involved. However, in an oper-
ating stope, a local pillar or face burst may be as destructive as a larger more distant
event slip located on a nearby fault. Techniques are required to identify mining layouts
that may be subject to each type of burst, and to develop preferred extraction layouts
and sequences to restrict burst frequency.
Two different types of support and reinforcement systems are required in hard rock
longwall mining and its variants. First, support is required for the hangingwall of the
mined-out void near and behind an advancing face. This support is usually described
as stope support. Second, support and reinforcement systems are required for the
access and transportation excavations, generally referred to as tunnels in South
African mines. Both types of support and reinforcement system may be required to
stabilize the rock mass under static loading conditions, thus reducing the risk of rock
falls and to alleviate the rockburst hazard under dynamic loading conditions.

3.5.2 Sub-level caving


Sub-level caving can be used for large orebodies, with steep dip and continuation at
depth. Adequate rock stability is required so that the sub-level drifts can be economi-
cally kept open, ideally with limited or occasional rockbolting only. The hangingwall
Underground hard rock mining 243

should be frangible and collapse as mining of the stopes progresses, to create the cave.
Significant surface subsidence occurs and cannot be readily stopped or controlled, so
as part of the mining permit, it must be allowed.
Caving requires a rockmass where both orebody and host rock fracture, under semi-con-
trollable conditions. As the mining removes rock without backfilling, the hangingwall
must keep caving consistently into the voids. Continued mining results in subsidence of
the surface where large sink holes may appear. Continuous caving is important to avoid
creation of cavities inside rock, where a sudden collapse could be harmful to mine instal-
lations by causing sudden, large stress increases on pillars and air blasts.
Sub-level caving methods allow the host rock to cave into the space created by blasting
and extraction of the minerals. This caving takes place in a controlled manner and can-
not be used in areas where subsidence is prohibited or there are significant aquifers. The
method is often used after a massive orebody has been mined as an openpit first. A network
of shafts, development drifts, crosscuts and haulage levels develops the mine complex.
The original application of sub-level caving (Figure 3.32) was in rock so weak that it
would collapse even in small headings where the support was recovered. Sub-level caving
has, however, been adapted to large ore bodies with steep dip and continuity at depth. The
hangingwall has to fracture and collapse continuously, forming the cave, and subsidence
of the ground surface above the orebody has to be tolerated. Sub-level footwall drifts must
be stable, requiring only occasional rockbolting. The key is that the orebody is blasted and
therefore consists of relatively small rocks, whereas the hangingwall caves under gravity
so the waste rock is much larger. This means that the ore can be preferentially loaded and
the loading stops when the large waste rock becomes evident in a drift.
The orebody is usually divided into sub-levels with relatively close vertical spacing at
approximately 8–15-m vertical intervals, depending on the plunge of the deposit. Each
sub-level is developed with a regular network of parallel drifts that penetrate the complete
ore section. Development to prepare sub-level caving stopes is extensive as compared to
other mining methods and mainly involves driving multiple headings to prepare sub-lev-
els. Ore is fragmented using blastholes drilled usually upward in fans from these sub-level
headings. Since the ore is blasted against the caved waste, explosive consumption tends to
be high. Ore is extracted selectively, with a LHD operating in the drill heading. This vehi-
cle transports the rocks to an ore pass where they are elevated to the surface.

Figure 3.32 Sub-level caving method (Atlas Copco, 1997)


1. Caved hanging wall 2. Blasting and loading 3. Drilling 4. Charging 5. Long-hole drilling
6. Developing of new sublevels 7. Sublevel 8. Footwall drift 9. Ore pass 10. Haulage level
244 Underground hard rock mining

Figure 3.33 General layout of sub-level caving method (Zhang & Wang, 2016)
1. Haulage drift 2. Ore pass 3.Sub-level drift 4. Drilling and loading crosscut 5. Slot drift
6. Slot raise 7. Finger long holes 8. Waste 9. Caved ore 10. Ramp entry

The general layout of sub-level caving method is shown in Figure 3.33.


On each sub-level, a system of the drives and crosscuts is developed following a
geometrically systematic layout. The drives are developed along the footwall and
hangingwall of the orebody and joined by a series of the parallel and regularly distrib-
uted crosscuts (Figure 3.34).

Figure 3.34 Sub-level caving method, generalized after Hamrin (2001)


Underground hard rock mining 245

Production of the sub-level caves is made by drilling the long fan (ring) blastholes into the
hangingwall from the sub-level crosscuts. The long blastholes are charged and blasted to
generate a controlled caving when hangingwall fractures and collapses following the cave.
The orebody can be mined by retreating from the hangingwall to footwall, which
is referred to as transverse sub-level caving (Figure 3.34), or conversely, retreating the
stopes along the strike of the orebody. The latter approach is used if the thickness of
the orebody is not suitable for transverse sub-level caving technique.
The rock mass must be stable enough to allow the sub-level drives to remain open
with limited support, usually local rockbolting and meshing only in unstable areas. It
general, the method can be used instead of sub-level open stoping (SLOS) if the rock
mass competence is insufficient for open stoping.
Transverse layouts are used for modern high production sub-level caving opera-
tions such as the Kiruna Mine, Sweden, and the Ridgeway Gold Mine, New South
Wales, Australia. Figure 3.35 (Queens University, 2012) is the generalized layout used
by Kiruna Mine in Sweden.

Figure 3.35 Sub-level iron ore mining at Kiruna Mine, Sweden (Queens University, 2012)
246 Underground hard rock mining

Figure 3.36 Mining layout for transverse sub-level caving (after Hamrin, 2001)

For wider orebodies, a transverse sub-level caving method may be used with the
production headings being driven across the orebody from footwall to hangingwall as
shown in the generalized mining layout in Figure 3.36.
A stability problem associated with sub-level caving is the brow stability in a sub-
level as shown in Figure 3.37.

3.5.3 Block caving (Cave Mining)


Block caving is the most complex and technically challenging underground mine
design, is the least flexible and highest risk (if the cave does not perform as designed).
It is also really the only mining method for high tonnage from a massive weak ore-
body. The production costs are relatively low due to the high tonnage and the closest
to those of openpit mines.
If the orebody is wide, steep, deep and frangible; block caving (Figure 3.38) would
be selected because the cost is lower than that of other suitable methods such as sub-
level caving or longhole open stoping. This method, sometimes called an upside-down
open-pit, is applied to large, massive orebodies in which areas of sufficient size can be
removed by progressively undercutting a slice, so that the rockmass above will cave
Underground hard rock mining 247

Figure 3.37 (a) A stable sub-level brow; (b) An unstable sub-level brow making it difficult to charge
the next row of blast holes (Hustrulid & Bullock, 2001)

naturally. The cave is controlled by the bulking of the breaking ore and the amount
that is removed from the drawpoints. It is applicable only to very large orebodies in
which the vertical dimension exceeds several 100m. The rather unique conditions limit
block caving applications to certain mineral deposits such as iron mineralization, low-
grade copper and molybdenum ores, and diamond kimberlite pipes.
Figures 3.38 and 3.39 show typical block caving mine layouts.
The main applications of cave mining are:

• Large massive orebodies


• Steep dip
• Large vertical extension
• Relatively weak and frangible orebody (readily cavable)
• Extreme surface subsidence can be tolerated

Figure 3.38 Block caving method (Atlas Copco, 1997)


1. Blastholes 2. Undercut level 3. Blastholes 4. Drawbell 5. Major apex
6. Production drift
248 Underground hard rock mining

Figure 3.39 Mechanized block caving at the El Teniente Mine, Chile (after Hamrin, 2001)

Block caving is based on gravity combined with internal rock stresses to frac-
ture and break the rock mass that has been undercut. Caving is induced by under-
cutting the block by blasting, destroying its ability to support the overlying rock.
Gravitational forces therefore act to fracture the block. Continued pressure and
movement created by drawing out the caved ore break the rock into smaller pieces
so they can pass out of the drawpoints where the ore is handled initially by LHD
loaders. This method is therefore different from all other underground mining
methods because, primary fragmentation of the ore collapses by natural gravi-
ty-induced mechanical processes. The elimination of primary drilling and blast-
ing has advantages in terms of orebody development requirements and other direct
costs of production. As fragmentation without drilling and blasting is uneven and
not totally reliable, a substantial amount of secondary blasting and breaking can
be expected at the drawpoints (typically around 30%). Therefore, the rock size, the
rate at which rock passes through the drawpoints and ensuring an even cave are
essential.
The method is carried out as follows:

• Each orebody block is completely undercut to the extent that the internal stresses
and gravity cause severe fracturing and then collapse, which initiates a cave (con-
trolled by the height between the undercut level and the rock (dilation, etc.). This
is governed by the hydraulic radius (discussed later).
• As the cave progresses upwards, the pressure on the drawpoints and the loose rocks
in the area, plus the movement caused by drawing off the rock cause autogenous
comminution (breaking the rock into manageable pieces typically say 1m³ or less
depending on the equipment and infrastructure design).
Underground hard rock mining 249

It can be very high tonnage and low cost but:

• Comprehensive predevelopment (drawpoints and undercut) is needed before pro-


duction starts, and build-up is initially slow.
• Poor fragmentations (creating huge boulders) are difficult to handle, can damage
the drawpoints and cause costly and potentially hazardous secondary blasting
costs.
• Drawpoint hang-ups can also create a non-uniform draw from the cave, leading to
higher localized stresses (static versus mobile soils) and even more stability issues.
Even draw is essential and must be carefully planned.

Where the ore block breaks up successfully and the extraction is carried out evenly
from all of the drawpoints, block caving becomes a low-cost, high-productivity method
with good ore recovery and moderate inflow of waste dilutions. Risks are high but the
result can be extremely favourable. This method is often used to convert an open-pit
operation into an underground mine where surface production can continue while the
underground infrastructure is prepared. In fact, the block caving method generates
production rates that can approach those of an open-pit (e.g., +100,000 tons per day).
Block caving is a large-scale production technique applicable to homogeneous ore-
bodies of large dimensions and most commonly steep to vertical orebodies. Block
caving is regarded as one of the most productive and lowest cost underground mining
methods and as such is often applied to low-grade orebodies. The term ‘block’ refers
to the entire volume of the ore which is prepared for excavation as a single block
(Figure 3.40).
Caving is induced by undercutting of a block. This is made by blasting the narrow ore
slice located underneath the block and removing the blasted ore (Figure 3.40). This cre-
ates a void beneath the block (undercut) and leaves the overlaying rock mass unsup-
ported. As a consequence of removal, this supports the rocks fracturing, induced by
the undercut blasting, which starts spreading through the entire block due to the grav-
ity force. The rock mass in a block breaks into smaller pieces which slowly move to
the drawpoint at the bottom of a stope (cave) where they are extracted through the
drawpoints.

Figure 3.40 Block caving method, generalized after Hamrin (2001)


250 Underground hard rock mining

The vertical height of a block caving stope can be as much as several hundred
metres. Therefore, even though the amount of development for the undercut level, the
production level and the haulage level is significant and an upfront cost, it is all that is
needed to remove a very large volume of ore relatively cheaply.
Application of the block caving technique is limited to large orebodies preferably
in a shape of a vertical cylinder and characterized by a favourable hydraulic radius.
Hydraulic radius (HR) is the term used to describe the point at which the under-
mined span of rock begins to cave under stress and gravity. It is:

HR = area/perimeter

Figure 3.41 shows the relationship of the Rock Mass Rating and hydraulic radius
based on many mines that use block caving.
The success of the method hinges on the understanding of the site-specific fracturing
process and capability of the broken ore to flow to the extraction points. In particular,
it is essential that the hangingwall of a stope is allowed to subside. Another condition
for the orebody to be amenable for block cave mining is the capability of the rock mass
to break into small pieces by the rock stress and gravity. All three conditions are rarely

Figure 3.41 Empirical data from mines using caving (Laubscher, 1994)
Note: MRMR considers rockmass strength, rockmass structure, in-situ stress, induced strengths, presence of
water and the hydraulic radius.
Underground hard rock mining 251

Figure 3.42 Block caving pattern (Hormazabal et al., 2018)

met; therefore, the method is used only for certain types of deposits; most commonly
these are copper and molybdenum porphyry deposits and diamond-bearing kimber-
lite pipes. The block caving is also used for mining some large iron-ore deposits.
Figure 3.42 shows a schematic picture of the alignment of the underground work-
ings in block caving relative to the stresses. Parallel drifts are developed through the
orebody on two main levels: the undercut level and the production level. In a later step,
the undercut initiation is conducted out of these drifts by drilling into the roof with
blasting. This weakens the groundmass directly. The breaking rock reaches the loading
drifts of the production level via draw bells. The production level is characterized by a
grid pattern of drifts consisting of the loading drifts and haulage crosscuts. This enables
a consistent distraction of the ore. The broken ore (muck) is carried by LHD from the
drawpoints to the collection points (ore-passes), which let the muck travel down to the
haulage level to be moved to the surface using a conveyor belt in a decline or hoisted in
a vertical shaft. Often a crusher is included to reduce the maximum size to one that will
not block the ore-passes and can be safely handled by the equipment suite. The haulage
level is connected to ramps or shafts, to convey the material out of the mine.
The undercut level is composed of numerous closely spaced parallel drifts (Figure 3.43)
connected by a main access, which runs perpendicular. The layout of the drifts is
adjusted so that the angle between the undercut drifts and the access could be enlarged
and matched to the curve radius of the mining equipment. The feasibility of this needs
further careful analysis and fitted to the unique geology of the orebody.
The inclination of the undercut drifts is such that water can be drained by gravity
to the shaft or ramp access.
Because the undercut drifts are abandoned once the cave starts, only short-duration
support is needed to protect the development crews.
The grid-like alignment of the drifts in the undercut and production levels is shown
in Figure 3.44.
The block caving method using a scraper is shown in Figure 3.45.
252 Underground hard rock mining

Figure 3.43 Undercut level (Hormazabal et al., 2018)

In block caving, mobilization of the ore into a caving medium is achieved without
recourse to drilling and blasting of the ore mass. Instead, the disintegration of the
ore (and the country rock) takes advantage of the natural pattern of fractures in the
medium, the stress distribution around the boundary of the cave domain, the limited
strength of the medium and the capacity of the gravitational field to displace unstable
blocks from the cave boundary. The method is therefore distinguished from all others,
in that primary fragmentation of the ore is accomplished by natural gravity-induced
mechanical processes.

3.6 Naturally supported mining methods


Naturally supported mining methods include:

• Room and pillar


• Sub-level stoping
• Longhole open-stoping (similar to sub-level stoping)

Figure 3.44 Production level, Finsch Diamond Mine, South Africa (Hustrulid & Bullock, 2001)
Underground hard rock mining 253

Figure 3.45 Block caving method using scraper, generalized (Zhang & Wang, 2016)
1. Haulage level 2. Crosscut 3. Boundary cut raise 4. Observation raise
5. Scraper drift 6. Ore pass 7. Scraper connecting roadway 8. Boundary cut drift
9. Observation crosscut

3.6.1 Hard rock room and pillar mining


Mining of steeply dipping or horizontal or near-horizontal orebodies is easier than at
“intermediate” dips (between say 30–55° from the horizontal). This is because with
steeply dipping orebodies, gravity is used for collecting the broken ore and waste. At
horizontal and near horizontal (<30° from the horizontal) orebody dips, room and
pillar mining is commonly used.
Room and pillar is an unsupported mining method that can be used for soft rock
(typically shallow coal seams) and hard rock orebodies (Figure 3.46). Examples of hard
rock deposits that are suitable to room and pillar are sedimentary deposits such as lime-
stone or sandstone containing lead, salt layers, phosphate, some base metal deposits,
limestone, magnesite and dolomite. This method uses a regular system of rooms in a
checker-board pattern, leaving regular square or rectangular pillars to support the over-
burden. Where the orebody height is significant (>6m), secondary benching can be used,
which is cheaper than the primary blasting because the bench has a second free face.
The dimensions of rooms and pillars depend upon factors such as the:

• Stability of the hangingwall, the ore and the footwall


• Thickness of the deposit
• The stresses both before and during mining
• Local and regional geology and hydrology
254 Underground hard rock mining

Figure 3.46 Room and pillar method (Atlas Copco, 1997)

The method can be rapidly adapted to changing conditions and allows selective
mining.
In this method, the ore is blasted (like in a tunnel) and loaded in the room and trans-
ported to a point where it will flow, either by gravity or mechanical means, to a central
gathering point to be taken out to the mine, via a decline (in trucks or on a conveyor
belt) or in skips up a vertical shaft.
Production from each room is discontinuous as it is blasted, and the tonnage per blast
is limited (maximum of say a 5m advance). The overall production can, however, be large
(around say 35,000 tpd) as numerous rooms can be blasted, mucked and supported daily.
Extraction is dependent on especially the depth and rockmass properties but can
vary from say 40–85%. A typical room and pillar layout for a thick deposit is shown
on Figure 3.46.
Where the orebody dips, the rooms can be developed on an apparent dip since
trackless equipment should be limited to a maximum of an 8–10° gradient. This is
shown in Figure 3.47.
Underground hard rock mining 255

Figure 3.47 Inclined room and pillar hard rock mining (Hamrin, 2001)

Rooms and pillars are commonly disposed in regular configurations to simplify


planning, design and operation, being designed with rectangular or square pillars
(square pillars being stronger than rectangular ones with the same cross-sectional
area).
Since personnel and equipment travel and work in the rooms (and intersections),
installed supports in the form of typically rockbolts or cablebolts are often needed.
Depending on the grade of the ore, secondary pillar extraction can be considered
on a retreat basis.
The main advantages of the room and pillar method are the:

• High degree of mining flexibility


• High degree of mechanization possible
• The costs are relatively low as most development is in the orebody itself, so
revenue-generating
• Ability to selectively mine
• Ability to mine (develop) multiple rooms at the same time
• Benching in high orebody heights that reduces costs

The main disadvantages are the:

• Hangingwall may need support that can be costly and reduce productivity
• Extraction becomes relatively low in weak and/or deep orebodies

With high orebodies in competent rock, slices can be taken from the bottom upwards,
reinforcing (laterally confining) the pillars with backfill, which also becomes the floor
for the subsequent slice. This means that smaller pillars can be used, thus increasing
the percentage extraction as shown in Figure 3.48.
256 Underground hard rock mining

Figure 3.48 High room and pillar mining with backfilling (De Souza, 2010)

3.6.2 Sub-level and open stoping (modified from Hamrin, 1982 &


Queens University, 2011b)
SLOS originated in the early 1900s in the iron ore mines of Michigan, US. It evolved
over the years from a method originally known as short hole bench and train mining.
This method is an alternate to sub-level caving and achieves less dilution but needs
a more competent rockmass. It is a relatively safe mining method as mining retreats
towards the solid away from the mined out area back towards the unmined and main
access (using typically ring blasting). Local ground support can be installed where
needed and where mine workers are or will be present, but the need for support can be
reduced especially if remote-controlled equipment is used.
There are numerous factors that affect the open stope length and width in sub-level
stoping, including:

• Principal stress directions and magnitudes


• Competence of the hangingwall
• Optimum drill pattern and available drilling equipment
• Orebody geometry
• Local and regional geology and hydrology
• Drilling drift layout

Proper design and dimensioning of each sub-level in this stoping method is critical to
ensure crew safety, optimum production and adequate equipment sizing. When dimen-
sioning the interval between subsequent sub-levels, both the height and width of the ore-
body must be taken into consideration. Typical sub-level intervals vary from 45m to 125m.
Sub-level stoping (illustrated in Figures 3.49 and 3.50) is used for mining mineral
deposits with steep orebodies and fairly regular boundaries, stable hangingwall and
footwall and competent orebody. This method requires extensive orebody develop-
ment with relatively high capital expenditures, but production costs are comparatively
low because most of the development is in ore. The thickness of the deposit between
the hangingwall and footwall usually varies from a few meters to tens of meters wide.
In this method, mining starts at the bottom of a level and proceeds upward. The ore-
body is vertically divided into levels, and between two levels, the stopes of convenient
Underground hard rock mining 257

Figure 3.49 Sub-level stoping method (Atlas Copco, 1997)

Figure 3.50 General layout of sub-level stoping method (Zhang & Wang, 2016)
1. Haulage drift 2. Ramp 3. Sub-level haulage drift 4. Ventilation drift 5. Sub-level
loading crosscut 6. Trough drift 7. Drilling access 8. Slot crosscut 9. Slot raise
10. Unloading crosscut 11. Ore pass 12. Crosscut 13. Crown pillar 14. Interval pillar
258 Underground hard rock mining

size are formed, hence the name sub-level stoping. Leaving a crown pillar at the top
of the stope safeguards the level above, while lower level is utilized as haulage level to
collect the mineralization from the stopes. The level developments, commonly in the
footwall, range from 50m to 150m, based on the vertical extent of the orebody and
the number of production openings that can be extracted at each level. Rib pillars are
often left between the stopes on the same elevation. Between the main levels, ramps
are usually driven for access and transportation. These ramps also give access to the
sub-levels, which are developed at regular intervals to remove blocks of ore. Dilution
with waste rock can occur if ore boundaries are irregular or if caving occurs, but usu-
ally almost all of the ore in the stope is recovered.
The ore is blasted from different sub-levels using longhole drilling, and the length
of the holes depends on the shape of the orebody and the sub-level spacing. The long-
holes generally do not exceed 25m because hole deviation and control can become
major problems beyond this length. The blasted ore reports to the drawpoints for
extraction. The loading can be carried out using remote-controlled LHD working in
the open stope, which reduces the amount of drift development in waste rock. Once
the stope is mined out, backfilling can be used. The backfill can be used as either a
filler, or if it is cemented, the pillars between the stopes can be extracted.
The stopes are developed using the sub-level drives which are prepared inside the
orebody between the main levels. The blast holes are drilled from the sub-level drives
distributed as a tight fan pattern covering the whole stope. The ore is broken by firing
the blast holes on the same cross-sections, usually referred to a firing ring. The SLOS
mining advances in a horizontal direction, usually along the strike of the orebody, by
drilling of the next firing ring, charging the blast holes and blasting them (Figure 3.50).
A reliable drawpoint system at the bottom of each stope is essential. Important
considerations include:

• Within the constraints of the orebody and hence the stope dimensions, the draw-
points must be designed with adequate spacing in order to guarantee uniform
drawdown and maximum recovery (based on the ellipsoid of draw).
• Correct floor preparation and roadway design to ensure excavation stability, easy
LHD manoeuvrability and effective water drainage.
• Space for the blasted ore because of the bulking.

The broken ore is removed by LHD from the drawpoints distributed along the stope
bottom (Figure 3.51). The recovery of the ore from the stopes is facilitated by trough-
shaped bottom of the stope (Figure 3.51). This technique creates large open voids,
particularly in the vertical direction. In order to prevent them from collapsing after
the ore is recovered, the stopes can be backfilled as already mentioned.
In conventional sub-level stoping where vertical rings or rows of holes are blasted,
it is necessary to have an initial slot to create a free face for blasting and to allow for
rock expansion. It can be done as a slot raise driven by:

• Conventional raising methods (slow and more hazardous)


• Alimak raising
• Raise boring
• Drop raising
• Crater blasting
Underground hard rock mining 259

Figure 3.51 LHD sub-level stoping method (Zhang & Wang, 2016)
1. Haulage drift 2. Crosscut 3. Raise for people and ventilation 4. Sub-level drilling
access 5. Trough 6. Crosscut for loading 7. Slot raise 8. Ventilation drift
9. Bottom pillar 10. Crown pillar 11. Interval pillar 12. Finger long holes

The slot usually extends from the extraction level to the back of the stope. Typically,
longhole slashing is used to expand the slot to the full stope width. A typical slot width
is 4–5m.
The method is used for steep dipping orebodies and requires stable rocks in the
hangingwall and footwall of the stope and also competent ore. Irregular shape of the
orebody and uneven contacts are undesirable because this can lead to excessive dilu-
tion of the ore. The same comments are related to internal waste distributed inside of
the orebody. The SLOS method doesn’t allow the separation of internal waste from
valuable material, hence everything inside the drilled pattern is recovered as ore.
Therefore, geologists reporting the ore reserves must make a correction for any inter-
nal and external dilution.
Carrying out proper undercutting (as shown in Figure 3.52) is critical to the success
of sub-level stope production due to its effects on the loading efficiency and produc-
tion blasting. To ensure stability, it should not be advanced by more than a few rings
before blasting.
Figure 3.53 shows the stoping sequence between interlevel drifts from the slot devel-
opment to the stope being completely mined out and Figure 3.54 shows an example of
ring drilling on a mine.
Generally, down hole drilling is more challenging than up hole drilling because drill
debris removal is more of an issue when drilling down holes.
260 Underground hard rock mining

Figure 3.52 Sub-level open stoping (Hamrin, 1982, 2001)

Figure 3.53 Longhole open stoping mining sequence between sub-levels (Goldfields, 2009)

In summary, the advantages of the method (over possibly block caving or more
commonly sub-level caving) are:
• Dilution may still occur (up to 20%) but total ore extraction from the stope is achievable.
Underground hard rock mining 261

Figure 3.54 Ring drilling pattern example from El Soldado Copper Mine, Chile (Hustrulid & Bullock, 2001)

• Pillars that are left in place can be removed later if financially viable once adjacent
stopes have been backfilled.
• Large-scale blasts reduce unit costs.
• Stopes may be pre-drilled before any blasting, to allow for larger and more effi-
cient blasts, assuming the blast holes once drilled remain stable.
• The initial development can be done in ore rather than in waste.
• The method lends itself to mechanization using large equipment.
• High productivity and efficiency can be achieved depending on the orebody.
• Repetitive mining operations and techniques help facilitate training and
safety.
• A relatively cheap mining method (more than cave mining).
• Relatively simple to ventilate.
• High recovery/extraction (75–90%).

The disadvantages include:

• Early stope production is low due to the lack of available drawpoints near
the slot, but it increases as stope extraction continues as more drawpoints are
exposed.
• Selective mining is difficult.
• Extensive early orebody development is needed with relatively high capital costs.
• The method is fairly inflexible.
262 Underground hard rock mining

• Drilling accuracy is important and should ideally deviate less than 2% on any
hole.
• Costs increase in narrower orebodies as lower production is achieved per blast.
• For orebodies with lower dips, dilution increases significantly.
• Noxious blast fumes can diffuse back into active stopes if secondary blasting is
required.

3.7 Artificially supported mining methods


Artificially supported mining methods consist of:

• Drift and fill mining


• Cut and fill
• Undercut and fill
• Shrinkage mining
• Vertical Crater Retreat (VCR)

3.7.1 Drift and fill mining


Drift and fill can be used for relatively flat dipping (typically less than 30° to the hori-
zontal) tabular orebodies whose grade (value) justifies the expense of the cemented
backfill as opposed to leaving pillars in a conventional room and pillar mine layout.
The backfill must have free-standing ability (so cemented) as it becomes the drift ribs
during secondary pillar extraction.
The method is a relatively low tonnage method and consists of regularly spaced
primary tunnels (drifts) developed in the orebody parallel to each other, at a span
that requires minimum support. Typically, the pillars left between the primary
drifts (tunnels) have the same width as the primary drifts. The height depends on
the height of the orebody, and if it is relatively high, the drifts can be cut in benches,
starting at the top (for easy access to install the roof support) and benching down.
The bench blasting is much cheaper than the primary excavation because it has two
free faces.
Figure 3.55 shows a primary drift 10m wide, using benching to achieve the full min-
ing height.
After each drift is mined to the stope limit, suitably designed barricades are erected
at both ends and the volume filled with a cemented backfill. When the backfill has
developed its target strength, the pillar between two backfilled drifts can be mined
(Figure 3.55).
It can be seen from Figure 3.56 that once the alternate primary drifts have been
filled with cemented fill, the secondary drift can be mined and either left unfilled or
filled with any (cheaper) free draining uncemented backfill. The planning of the sec-
ondary pillar extraction must be carefully considered as the cemented backfill in the
primary drifts must have an adequate strength to free-stand and not be significantly
affected by the secondary extraction blasting.
The method can be modified to mine suitable, more massive orebodies where for
some reason a more economic method like longhole open stoping cannot be used.
Figure 3.57 shows an example of such a hybrid method.
Underground hard rock mining 263

Figure 3.55 A drift being developed and benched


Cooke 3 Shaft, Randfontein Estates Gold Mine, South Africa, now mined out and closed

Figure 3.56 Section showing a simplified drift and fill mining sequence

3.7.2 Cut and fill mining


Cut and fill mining is used typically under the following conditions:

• Steep dipping (typically +55° from the horizontal) tabular (vein) type deposits.
• Relatively narrow width.
264 Underground hard rock mining

Figure 3.57 A modified drift and fill and undercut and fill (Gertsch & Bullock, 1998)

• Relatively stable orebody as the orebody is the roof of the stope as it progresses.
• The orebody can have a varying width and grade, as selective mining is
possible.

The orebody must have a relatively high value as the mining method produces a
low tonnage. As mentioned, selective mining is possible too (where unpaid areas
can be left unmined). The width of the orebody can be variable too as it is easy
to follow the width of the orebody based on the previous lower cut and the local
geology.
In cut and fill stoping, the orebody is extracted in horizontal slices beginning at the
very bottom of a stope and advancing upwards towards the top of the stope or surface.
Ramps (inclined tunnels) are excavated to connect the surface to the underground
orebody and to access the individual stopes. The ramps are developed at a safe dis-
tance from the orebody so that they are not damaged by the mining-induced stresses.
Drifts (sometimes called crosscuts) are excavated from the ramp to the orebody. An
outline of a cut and fill mine is shown on Figure 3.58.
The slices are drilled using a mechanized drill rig (or hand-held), blasted by charg-
ing the drill holes with explosives, and ore is removed by using dump trucks or LHD
vehicles. The ore is dumped into an ore pass (a small steeply inclined tunnel) where
ore is transported to a lower elevation in the mine. The ore is picked up at the other
end of the ore pass by a scoop tram or LHD to be transported out of the mine usually
through a ramp (inclined tunnel). Once a slice is completely mined out, the empty
space is partially backfilled hydraulically. The backfill material used can be a mix-
ture of sand and rocks, waste rock sometimes with cement, or dewatered mill tail-
ings (rejected low-grade ore from processing, usually fine and sandy). The backfill
underground serves to keep the mine side walls stable and also as the floor for mining
the next slice. Mining continues upwards towards the surface until the orebody is
depleted. The main backfill requirement is that it must be free draining as it acts as
Underground hard rock mining 265

Figure 3.58 Outline of a cut and fill mine with ramp access (Atlas Copco, 1997)

the floor for subsequent cuts so must support equipment and men. An example of a cut
and fill layout is shown in Figure 3.59.
The support of each subsequent cut (or lift) can be easily determined using dam-
age mapping on the cut below and extrapolating it to the next cut above. Less stable
areas can then be anticipated and appropriate support installed as they are mined, or
pre-supported using spiling bolts.
After the stope cut is complete to the stope length and is totally mucked clean (the
blasted ore removed), the mined-out space is backfilled to form a working platform so
the next cut (slice) can be mined. A layout with the ramp access that is raised as cuts
are taken is shown on Figure 3.60.
Development for cut and fill mining in a trackless environment includes a footwall
haulage drive along the orebody at the main level, undercut of the stope provided with
drains for the hydraulic backfill, a spiral ramp excavated in the footwall with access
turn-outs to the stopes and a raise from the stope to the level above for ventilation and
fill transport.

3.7.2.1 Overhand stoping


Overhand stoping is used with cut and fill mining, with both dry rock and hydrau-
lic sand as backfill material. Overhand means that the ore is drilled from below by
266 Underground hard rock mining

Figure 3.59 Example of a cut and fill stope with rail ore transport (Hamrin, 2001)

Figure 3.60 A
 cut and fill stope showing the ramp access that is blasted and raised after each cut
using typically waste rock (Atlas Copco, 1997)
Underground hard rock mining 267

blasting a slice 3–4m thick. This allows the complete stope area to be drilled and the
blasting of the full stope without interruptions. The “upper” holes are drilled with
simple drill rigs.

3.7.2.2 Up-hole stoping


Up-hole drilling and blasting leaves a rough rock surface for the roof; after mucking
out, its height will be about 7m. Before miners are allowed to enter the area, the roof
must be secured by trimming the roof contours with smooth-blasting and subsequent
scaling of the loose rock. This is usually done by miners using hand-held rock drills
working from the muckpile.

3.7.2.3 In-front stoping


When using in-front stoping, trackless equipment is used for the ore production.
Sand tailings are used for backfill and distributed in the underground stopes via
plastic pipes if possible as they are easier to transport and install and cheaper than
steel. The stopes are almost completely filled, creating a surface sufficiently hard to
be traversed by rubber-tyred equipment. The stope production is completely mech-
anized with drifting jumbos and LHD vehicles. The stope face is about 5m high
across the stope with a 0.5m open slot left beneath it to facilitate the horizontal
row-by-row blasting, which causes bulking, hence the need for the gap (Figure 3.61).
Five-meter-long horizontal holes are drilled in the face, and ore is blasted against
the open bottom slot.
Figure 3.62 shows a small narrow hand-drilled vein cut and fill stope.
Scraper cleaning especially in very narrow (hand-got) deposits is still quite common
and effective for cleaning after each blast.

3.7.2.4 Backfill
Backfilling is the key to the method, and after each cut (slice) is mined to the limits of
the stope panel, it is backfilled. The backfill acts as the working floor for the next slice.
The requirements for the fill are that it must be “free draining” so men and equipment
can work safely and effectively. It does not need to be cemented as it does not need to
be “free standing”.
Once a whole cut is mined, it must be backfilled before the next cut above can be
mined. Various backfill types can be considered, such as:

• Rockfill (waste rock)


• Classified slurry tailings
• Uncemented paste fill (probably not cost effective)

The rockfill is placed either using a truck or LHD, ensuring it is placed in good con-
tact with the ribs (cut sidewalls).
Classified slurry tailings are placed behind a simple barricade that can be made
of wall of waste rock or tailings. The hydraulic pressure against the barricade is
low due to the low cut height (typically less than 5m). Depending on the length of
268 Underground hard rock mining

Figure 3.61 D
 rilling a new face with the “free surface” created by leaving a gap between the backfill
and the next cut (modified from Gertsch & Bullock, 1998)

Figure 3.62 Small cut and fill stope at the Historical Lucky Friday Mine, US
Reproduced with kind permission from Hecla Ltd.
Underground hard rock mining 269

the stope and the access position to the stope, the fill may need to be introduced in
multiple sections that may each need a barricade. A facility must be made available
to remove excess water runoff, so the fill consolidates and stabilizes to be able to
handle traffic.
Unconsolidated fill is placed in the heading using either a truck, LHD or blasted
from the cut above. The fill can either be jammed against the face to ensure tight fill
against the back or a small void (0.5–1m) can be purposely left behind to act as a free
face when advancing the next overcut (see Figure 3.62).
Advantages:

• Low-strength backfill or unconsolidated fill is adequate


• Low wait time to allow backfill to cure
• Use of run of mine waste as backfill reduces cost
• Ability to be selective

Disadvantages:

• Potentially more up-front development to access the bottom of the zone


• High ground support consumption requirements
• Mining through installed ground support
• Labour intensive
• Higher mining cost than bulk mining methods
• Sequence dependent on fill placement

Risks:

• Subsidence of unconsolidated backfill


• Dilution from mucking the unconsolidated fill in the floor, as some of the blasted
ore can get lodged into the fill

The backfill can be transported to the stope by:

• Pipelines (most common)


• Trucks
• Conveyors
• A combination

3.7.2.5 Men and materials access way


A cut and fill stope can be either accessed the whole time within a full stope panel by
using a ramp access, which is raised after each cut is taken, or it can be a “captive stope”
with access for equipment only at the first and last cut in a whole stope (Figure 3.63). In
this option, access can be via shuttered access ways in the backfill or the rock (created
by a raise). Whilst this is less costly as the access ramp does not need to be blasted,
raised and “re-levelled” using fill, if there are major equipment breakdowns, the repair
is often slow and difficult due to access constraints.
270 Underground hard rock mining

Figure 3.63 A captive cut and fill stope (Queens University, 2011a)

This is also a common access method in shrinkage mining too.


Figure 3.64 shows a captive cut and fill stope with scraper cleaning.
Support is installed as needed and cable bolts are commonly used as they are less of
an obstruction during the mucking (cleaning). The exposed cable support can be cut
off using an angle grinder or cutting torch if needed.
Figure 3.65 shows a hand-drilled cut and fill stope that uses cablebolts for
support.
Damage mapping is useful and involves identifying weaker localized strata where
more support is needed and projected to the slice above so the required support can
be “pre-planned” and installed in advance.

Figure 3.64 A captive cut and fill stope with scraper cleaning (Gertsch & Bullock, 1998)
Underground hard rock mining 271

Figure 3.65 Sequence of cablebolting (if required) in a cut and fill stope (Hoek et al., 1993)

3.7.3 Undercut and fill mining

3.7.3.1 The selection between cut and fill and undercut and fill
The selection of cut and fill or undercut and fill depends mainly on the strength of the
orebody and whether it will be stable as the roof of a cut. Clearly, cut and fill is the
cheaper mining method because in undercut and fill, the backfill is the engineered
roof of the cut so it needs to be cemented as it becomes the roof of the next lower cut.
Figure 3.66 shows how the selection and mining width (cut width) is mainly dependent
on the ground conditions.
The main differences between cut and fill and undercut and fill are shown in
Figure 3.67.
This method involves mining sequentially in horizontal lifts beginning at the
top of a stope and working downwards. A binder is added to the backfill to create
a stable roof under which to work. Typically, cement is used as a binder. The lower
portion of the fill pour will often have a higher cement content (~14%) and the
upper part of the pour would have a lower cement content (3–5%). Undercut and
fill can also be mechanized or hand-got with continuous ramp access or captive
conditions, the same as for cut and fill. It is less productive and costly so is only
272 Underground hard rock mining

Figure 3.66 Mining selection method for thin steeply dipping tabular orebodies

applicable for rich deposits (typically gold and silver). Examples of layouts are
shown in Figures 3.68 and 3.69.
Once a cut is complete to the stope boundaries, it is typically pre-supported as
it will become the roof of the next lower cut so needs to be correctly and consist-
ently engineered. The support obviously needs to be installed on the floor of the
completed cut. Commonly, rockbolts with large plates are placed on the floor at the
desired spacing. Timber or steel straps could be added followed by screen (mesh)
and some timber or steel plates at the end of each rockbolt as shown in Figures 3.70
and 3.71.
Once completely backfilled, the rockbolts behave like soil nails such as those com-
monly used in underground civil construction projects to stabilize soils or clays.

Figure 3.67 The difference between cut and fill and undercut and fill (Williams et al., 2007)
Underground hard rock mining 273

Figure 3.68 Typical undercut and fill mine section (Gertsch & Bullock, 1998)

Figure 3.69 A scraper cleaned undercut and fill captive stope (Pugh & Rasmussen, 1998)
274 Underground hard rock mining

Figure 3.70 Pre-backfilling preparation of a completed cut (Gertsch & Bullock, 1998)

Figure 3.71a shows the mesh and rockbolts with plates attached to both sides on the floor
(footwall) of the undercut before it is filled. Figure 3.71b shows the start of the cemented
backfilling. Often, the first placement of cemented backfill is stronger (i.e., has a higher
cementitious content) than the rest as it is the immediate roof for the next undercut.
According to Raffaldi et al. (2018), a unique challenge to underhand cut and fill
mining at Lucky Friday was dealing with horizontal stress induced in the cemented
backfill as a result of stope wall closure. During the undercutting, the stope walls
converge in response to the high horizontal ground stresses and compress the backfill.
These horizontal loads cause crushing and extensional fracturing near the fill surface
as shown in Figure 3.72.

3.7.4 Shrinkage mining


Shrinkage stoping is a generic term used in mining to describe the process of mining
upwards from a lower to a higher horizon, leaving broken rock in the excavation as
the working floor and the support of the sidewalls of the stope. The broken rock acts
as a working platform and helps stabilize the excavation by supporting the waste rock
walls. The method may be used for ore mining in shrinkage stopes, for rising, and for
underground construction projects where excavations of considerable vertical height
may be needed, like ore and waste bins, crusher rooms, penstocks and tailrace tunnels.
The requirements for shrinkage stoping are:

• Steeply dipping (more than the angle of repose of the rock)


• Stable ore (not subject to deterioration/weathering after blasting with time)
Underground hard rock mining 275

Figure 3.71 Backfilling an undercut and fill stope at Historical Lucky Friday Mine
Reproduced with kind permission from Hecla Ltd.

• Relatively stable footwall and hangingwall


• Regular orebody boundaries
• Relatively stable orebody

As blasted rock takes up more volume than rock in situ, some of the broken material
must be removed in a periodic and planned manner to maintain the requisite relation-
ship among the back (or roof) of the excavation and the level of the broken material
(floor) in the excavation. This is attained by drawing the blasted material through draw-
points on the lower level, which is constructed prior to start of shrinkage stoping. Access
to the space among the broken material and the back of the excavation ought to be
maintained for access of men and materials and for ventilation. This type of access is
usually provided by previously installed raises, normally equipped with ladder ways.
When correctly planned and executed, shrinkage mining is a very effective tech-
nique for ore mining and underground construction. It is used where the hangingwall
and footwall of the excavation are strong enough to be self-supporting, although some
artificial support like rock and cable bolts might be installed as shrinkage progresses.
276 Underground hard rock mining

Figure 3.72 S urface extension fracturing in the cemented paste backfill beam due to horizontal closure
at Historical Lucky Friday Mine
Reproduced with kind permission from Hecla Ltd.

Where the technique is used for ore mining, careful planning and scheduling are req-
uisite to ensure consistency of ore grade and manufacture tonnage.
Shrinkage stoping is used in steeply dipping, relatively narrow orebodies with fairly
regular boundaries. Ore and waste should be sturdy, and the ore must not be affected
by storage in the stope.
From 30 to 40% of the broken ore is taken from the bottom of the stope, and the ore
in the slice is blasted down, restoring the volume withdrawn (i.e., the amount of bulking
from blasting). The miners then re-enter the stope and work off the recently blasted ore.
Shrinkage stoping is rather hard to mechanize; in addition, an important period
may elapse between the commencement of mining in the stope and the last withdrawal
of all the broken ore, once the stope is completed.
Consecutive horizontal slices of ore, roughly about 3m high, are taken along the length
of a stope, in a manner similar to cut and fill mining. The ore is taken out from the stope
regularly from each closely spaced drawpoint at the underneath horizon spaced typi-
cally every 7.5m along the strike. Adequate ore is left in place to give a floor from which
to work when taking the next cut. This needs considerable planning and coordination.
When the ore is blasted, it fills a space about 1.3–1.5 times the size of the space it
filled as the solid un-blasted mass. This is known as bulking or swell and is a vital
factor in determining how much ore to draw from the bottom of the stope in order to
maintain adequate working room. The broken ore is drawn down from chutes under-
neath, therefore “shrinking” the volume of broken ore in the stope.
The procedure is continued upward until the stope either reaches the next level or is
stopped at some predetermined elevation. Horizontal crown pillars are left at the back
(at the top) of the stope.
Underground hard rock mining 277

Shrinkage stoping relies on gravity to keep the broken ore moving to the draw-
points, so it works only in steeply dipping orebodies. The sidewall rocks (the hang-
ingwall and footwall of the reef) must be strong and capable. The orebody ought to
be wide enough for efficient mining, so usually 2m or more. Examples of layouts are
shown in Figures 3.73 and 3.74.
It is important to draw evenly out of the stope to create a safe and solid working
floor with no voids in it.

Figure 3.73 Shrinkage stope layout (Hamrin, 2001)


278 Underground hard rock mining

Figure 3.74 Section and plan of a hand-got shrinkage stope (Michaud, 2016)

Figure 3.74 shows the layout for a hand-got shrinkage stope.


The blast round design is basically a wedge cut to create the free face and is rela-
tively simple as shown in Figure 3.75.

3.7.5 Vertical Crater Retreat (VCR)


The VCR mining method (Hustrulid, 1982) was developed by the International Nickel
Corporation (INCO), now owned by Vale out of Brazil. It is based on C.W. Livingston’s
crater blasting theories. It was first used in 1974, at the Levack Mine located in the
Sudbury Basin’s North Range, Canada, and it provided productivity benefits almost
immediately. Mining productivity, measured in pounds of nickel and copper produced
per manshift, improved by approximately 80% in mines in the Sudbury Basin between
1980 and 1990, after their conversion to VCR mining. The mining method soon spread
worldwide, being adopted for use for the first time in the US in 1977 at the Homestake
Mine (Queens University, 2014).
VCR is used at the deposits that have competent steeply dipping ore hosted in com-
petent wall rocks (Figure 3.76). The ground conditions in general are similar to that of
the SLOS method; however, the VCR is technically simpler and allows typically larger
production rates at lower production costs. The orebody must be competent enough
that the whole width can be excavated to form the drilling drifts.
The method is based on the crater blasting which is made by firing the large-diameter
blast holes (typically 15–20 cm) drilled downward from the overcut developed on the top
of a stope (Figure 3.76). Part of the blasted ore is left in the stope and acts as a temporary
support to the stope sides (the orebody hangingwall and footwall in a manner similar to
shrinkage stoping). The blast holes are charged by explosives placed into the short sec-
tions of the holes, which is referred to as crater charge. The explosives are positioned in
each hole at the same distance above the open surface (Figure 3.76). The blasting loosens
the ore slice, creating a crater which is vertically retreated. Ore is continuously extracted
Underground hard rock mining 279

Figure 3.75 A typical blasting round for a shrinkage stope (Gertsch & Bullock, 1998)

from the drawpoints together with rigorous recording of the blasting progress in each
hole. Since no men are in the stope, it is much safer than shrinkage mining.
According to Queens University (2014), the advantages and disadvantages of the
method compared to SLOS and shrinkage mining are as follows:

• Advantages:
• Safety: Miners are working in drifts that are adequately ventilated and have
fully supported roofs. Furthermore, no workers are required to work inside
the stope, minimizing the risk of unexpected injuries. VCR mining also allows
for the use of automated machinery, in which case workers are not at risk of
equipment-related injuries.
• Good ore recoveries: Continuous mucking from the drawpoints can take
place after blasting. Furthermore, VCR mining can be used with a high degree
of mechanization, generating a high level of productivity.
• Cost: Once the pre-mining development is in place, mining has a low operat-
ing cost, as it is a bulk mining technique and employees are not required to
manually operate the mucking machinery.
280 Underground hard rock mining

Figure 3.76 Vertical crater retreat mining method (Hamrin, 2001)

• Wall support (orebody hangingwall and footwall): VCR stoping shares


some great features with sub-level open and shrinkage stoping. Good wall
support is offered during the VCR stoping phase, using the shrinkage
technique.
• Disadvantages:
• Dilution of the ore can result if waste rock is less than competent in strength,
or improper blasting techniques are carried out. In addition, sorting is not
possible.
• Risk of drawpoint blockage: If improper blasting techniques are carried out,
large rocks can get lodged in the drawpoint and arrest the movement of mate-
rial through them. This results in lost production.
• A large capital investment is required to establish the essential drift infra-
structure required for proper VCR mining to take place. Furthermore, the
acquisition of equipment is necessary before mining can occur.
• Subsidence of overlying stope zones can be a problem as open stope
expanses are left after mucking. This can be avoided by using a back-
fill, which obviously comes at an additional cost, but is good for the
environment.
Another general layout of this method is shown in Figure 3.77.
Figure 3.78 shows sections through a VCR stope where the first blast of the stope
has been taken.
Underground hard rock mining 281

Figure 3.77 General layout of vertical crater retreat mining method (Zhang & Wang, 2016)
1. Bolting 2. Drill drift 3. Haulage drive 4. The third blasting layer 5. The second blasting
layer 6. Crater charge 7. The first blasting layer 8. Undercut level 9. Loading
crosscut 10. Drawpoints bell

3.8 Drilling and blasting

3.8.1 I ntroduction to blasting (www.science.jrank.org/pages/2634/


Explosives-History.html)

3.8.1.1 The history of explosives


The first chemical explosive was gunpowder, or black powder, a mixture of char-
coal, sulphur and potassium nitrate (also known as saltpeter). It was invented by the
Chinese over 1,000 years ago and used mainly for fireworks at celebrations. It was

Figure 3.78 Sections through a VCR stope (Gertsch & Bullock, 1998)
282 Underground hard rock mining

first used as a weapon of war by the Europeans for firing cannons, etc., who probably
learnt about it from traders in China.
The English and the Germans manufactured gunpowder from the early 1300s. It
remained the only explosive for about 300 years, until in 1628 another explosive called
fulminating gold was discovered. Fulminating gold was the first high explosive known
to man and was first noted in western alchemy as early as 1585. Sebald Schwaerzer
was the first to isolate this compound and comment on its characteristics in his book
Chrysopoeia Schwaertzeriana.
Gunpowder changed the lives of both civilians and soldiers in every Western
country that experienced its use. Eastern nations like China and Japan rejected the
widespread use of gunpowder in warfare until the nineteenth century. Armies and
navies who learned to use it first, for example, the rebellious Czech Taborites fight-
ing the Germans in 1420 and the English Navy fighting the Spanish in 1587, relied on
it heavily for their respective victories. These victories, however, quickly forced their
opponents to learn to also use gunpowder effectively. This changed the way wars
were fought. This meant that taking refuge behind high thick walls was no longer
necessarily safe. Gunpowder could be used to fire projectiles (usually cannonballs)
against the same position in the wall and thereby cause it to collapse. Large and
united regions with larger armies became more important and powerful than local
despots and rulers, and war became more deadly, due in large part to the use of
gunpowder. It was not until the seventeenth century that Europeans began using
explosives in peacetime to blast rocks in mines and clear obstacles such as boulders
and large trees.
The first recorded use of an explosive in a mine was in Saxony, France, in 1627. The
widespread application in mines only occurred after the development of the safety
fuse in 1831 by Bickford.
In 1846, Italian chemist Ascanio Sobrero (1812–1888) invented the first modern
explosive, nitro-glycerine, by treating glycerine with nitric and sulfuric acids. Sobrero’s
discovery was, unfortunately often too unstable to be used safely. Nitro-glycerine
readily explodes if suddenly bumped or knocked. This inspired Swedish inven-
tor Alfred Nobel (1833–1896) in 1862 to seek a safe way to package nitro-glycerine.
In the mid-1860s, he succeeded in mixing it with an inert absorbent material and called
his invention dynamite.
Since then, other chemical explosives have been discovered, the most common of
these are chemical compounds that contain nitrogen such as azides (compounds con-
taining N3), nitrates and other nitro-compounds.
Significant dates in the early development of explosives and accessories are:

1846 – Nobel develops nitro-cotton and nitro-glycerine.


1867 – Nobel invents dynamite and the mercury fulminate detonator that made
the use of dynamite safe. Dynamite replaced gunpowder as the most widely
used explosive (aside from military uses of gunpowder).
1875 – Nobel invents blasting gelatine. Nobel continued experimenting with
explosives, and in 1875, invented a gelatinous dynamite, an explosive jelly. It
was more powerful and even a little safer than the dynamite he had invented
9 years earlier. The addition of ammonium nitrate to dynamite further
decreased the chances of accidental explosions. It also made it cheaper to
manufacture.
Underground hard rock mining 283

1879 – Nobel invents ammonium nitrate and nitro-glycerine, making a whole


range of low-strength explosives. These and other inventions made Nobel
very wealthy. Although the explosives he developed and manufactured were
used for peaceful purposes, they also greatly increased the destructiveness of
warfare. When he died, Nobel used the fortune he made from dynamite and
other inventions to establish the Nobel prizes, which were originally awarded
for significant accomplishment in the areas of medicine, chemistry, physics
and peace.
From 1955 – ANFO (ammonium nitrate and fuel oil) was designed and later the
range of slurry explosives.
1960s – NONEL is a shock tube detonator (“Non-Electrical”) that has an explo-
sive compound deposited on the inside of the plastic tube. It was invented by
Per-Anders Perrson from Dyno-Nobel.
1990s – Development of the milli-second electronic delay detonator (EDD).

3.8.1.2 Definition and components of an explosive


An explosive is generally defined as a solid or liquid substance, or mixture of sub-
stances, which on application of a suitable stimulus to a small portion of the mass, is
converted in a very short time into a more stable compound that is largely or entirely
gaseous, with the consequent development of heat and very high pressure.
The basic components of any common commercial explosive are:

• Oxygen supplier
• Fuel
• Sensitizer
• Stabilizer

There are two major components in an explosive: fuel and oxidizers (oxygen suppliers).
These two components are balanced and used according to what type of detonation
is needed. This chemical composition is the backbone of the explosive and has been
improved upon over time. Products such as ANFO (Ammonium Nitrate and Fuel
Oil), also known as ANFEX (Ammonium Nitrate Fuel EXplosive) have a high level
of insensitivity, resulting in the need for boosters because of the material’s inability to
detonate effectively via a shock initiation.

OXYGEN SUPPLIERS (OXIDIZERS)

Examples of oxidizers are ammonium nitrate, calcium nitrate, sodium nitrate and
black powder. Oxidizers are compounds which are capable of reacting with and oxi-
dizing other materials. They “give” oxygen molecules to other materials, such as fuels,
in which they burn and create gases and pressure.

FUELS

Examples of fuels are fuel oil, carbon, aluminium, tri-nitro toluene (TNT), research
development explosive (RDX) and smokeless powder. Fuel oil is one of two components
in ANFO, the other being ammonium nitrate. The fuel accounts for approximately 6%
of the chemical composition. In underground applications, the percentage of fuel oil in
a product such as ANFO is precisely calculated to reduce the production of fumes.
284 Underground hard rock mining

Explosives are divided into two categories depending on the velocity of the explosive
reaction:

• Detonating explosive – the explosive reaction speed is more than the speed of
sound
• Deflagrating explosive – the explosive reaction speed is less than the speed of sound
(e.g., gunpowder)

3.8.1.3 How explosives work and rock fragmentation


The explosive action causes three zones as loosely defined and shown in Figure 3.79:

• Zone 1 – This represents the immediate area around the blast hole, which is com-
pletely pulverized and shattered. The compressive stresses caused by the explosive
shock wave far exceed the compressive strength of the rock.
• Zone 2 – This zone is characterized by radial cracks extending out from Zone 1.
The cracks are caused by the tangential component of the stress wave radiating
out from the hole.
• Zone 3 – Further damage is caused in this area due to the gas pressure that causes
certain of the radial cracks to extend even further.

Under the influence of the wave, the rock particles are moved radially outwards,
resulting in a tensile stress in the tangential direction (and rock is weak in tension).
This stress wave is large enough to cause radial fractures.
Aside from the immediate formation of heat and fumes, four main effects of det-
onation are rock fragmentation, rock displacement, ground vibration and air blast.
These four effects are of particular importance in mining. The fracturing of rock, fly
rock, vibration of ground and shockwaves can cause serious problems in mining and
the surrounding community.

Figure 3.79 The zones created by an explosion


Underground hard rock mining 285

Figure 3.80 An underground blasting round


Photo courtesy of Dyno Nobel

While fragmentation is the primary wanted result in blasting, a balance is sought


to prevent pulverization or lack of fracturing. Different applications will require
varying degrees of rock breakage. The type, amount and placement of explo-
sives all affect fragmentation. If the breakage is too large, secondary blasting is
needed, raising production cost. In addition, if the muck pile is too large, bucket
loads will be decreased in overall volume, decreasing production and increasing
cycle times. If the rock is pulverized, material fines processing goes up, increasing
the cost.
Figure 3.80 shows an underground tunnel being blasted (after Dyno Nobel).
Tamping and stemming are also important to optimize the blasting efficiency.
Tamping generally means the consolidation of stemming and blasting materials in a shot
hole.
Stemming means filling the shot hole (near the entrance) with an inert material.
This helps to provide that little extra confinement to the blasting medium to get
more effective fragmentation. Inert material means it does not react during the
blast.
In order to increase the efficiency of explosive comminution, the borehole pressure
must be maximized and pressure losses minimized. The majority of these pressure
losses occur from premature borehole venting and through weak rock layers intersect-
ing the borehole. With the use of proper stemming material and amount of stemming,
these pressure losses can be minimized, increasing the efficiency of explosive commi-
nution (Green Energy & Technology, Springer). Effective stemming also helps reduce
overbreak.
Generally, wet blast holes require more stemming than dry blast holes. Figure 3.81
shows how a vertical blast hole is stemmed and Figure 3.82 shows the stemming in a
horizontal blast hole.
286 Underground hard rock mining

Figure 3.81 Stemming in a typical up-blast hole (Zhang, 2008)


A, B and C are detonators and the shaded area from A to the tunnel perimeter would be stemming

3.8.2 Optimizing blasting


The efficiency of blasting must always be continually reviewed and improved. Typical
considerations are given in Figure 3.83.
Successful drilling and blasting depend on matching both with each other, and
also the rock conditions. The 3 main parameters needed to optimize explosive perfor-
mance are given on Figure 3.84.

Figure 3.82 Stemming in a typical underground blast hole


Underground hard rock mining 287

Figure 3.83 Blast considerations

3.8.2.1 Blasting in coal mines (and other fiery mines)


In coal mines, where methane commonly occurs and a mixture of 5–14% methane
forms an explosive mixture in the air, the detonation of explosives could cause a sec-
ondary and serious methane explosion. Special or “permitted” explosives are there-
fore used that minimize the danger of a methane explosion. Permitted explosives have
the following properties:

• The reaction temperature is reduced to a safer level.


• No constituents in the explosive continue to burn after the main reaction has been
completed.

3.8.2.2 Blasting principles


Site-specific geology must always be considered when planning any blasting activity.
Nearly all of the blasting problems are caused by geological features. Explosive energy

Figure 3.84 Explosive performance


288 Underground hard rock mining

level and explosive distribution must be matched to geological conditions in order for
a blast to be successful. The initiation timing sequence must be designed to work
with the specific rockmass response to the explosive loading. The blast engineer or
responsible person cannot make any necessary adjustments without some informa-
tion and understanding of the geology in the blast area. Modifications are made
to the standard blast design based on observation of the blasted area and on past
experience.
Blasting is an important component of the rock breakage cycle in hard rock mines.
Blasting is used to fragment, pulverize, move or otherwise displace rock in order to
produce a pile of fragmented rock that can be gathered using the available equipment.
Blasting is one of the four key components of a typical metal/non-metal mining pro-
duction cycle that consists of:

• Drilling
• Blasting
• Loading
• Making safe and supporting

Figure 3.85 shows how the explosion front propagates through a charged hole.

3.8.2.3 Explosive properties


There are several major explosive properties which define and quantify a detonation:

• Detonation velocity
• Explosive density
• Detonation pressure
• Explosive energy
• Water resistance of the explosive
• Sensitivity
• Sensitiveness
• Blasting fumes
• Flammability

Figure 3.85 Explosion detonation diagram, showing the Chapman-Jouget Plane (Das Sharma, 2012)
Underground hard rock mining 289

These properties are used to help decide the type of explosive needed, the strength
of explosive needed, any resistance to water requirement and how it is loaded into the
blast holes for a specific application.

DETONATION VELOCITY

Detonation velocity is the velocity at which a shockwave front travels through an


explosive upon detonation. This variable is used in calculating detonation pressure.
The detonation velocity should always be higher than the speed of sound of the mate-
rial being blasted. Typical explosive velocities range from 3,000 to 5,500 m/s. Variables
such as charge diameter, temperature, confinement method and primer application
affect the final detonation velocity of an explosive too.
Charge diameter plays an important role in detonation. The larger (wider) an explo-
sive’s reaction zone, the higher the dependence on diameter. There exists a “critical
diameter” in which a material will fail to explode effectively if it falls below this number.
Temperature is a variable that must be taken into consideration for commercial
blasting applications. In areas where freezing is common, water-based explosives will
need to be adjusted accordingly. Solid explosives such as ANFO are commonly used
in areas where low temperature is a concern.
The confinement method is of vital importance in any explosion. This is evident in
matching blast hole diameters to charges and within the explosive itself. Confinement
creates higher pressure which is desirable for high detonation velocities. Typically, the
larger the diameter of the bore hole, the less reliant the charge is on confinement (i.e.,
the confinement has diminishing returns as diameter increases).
Lastly, a primer can also play an important role in detonation. The misuse of
priming, however, can result in slower velocities, failed detonation and/or deflagra-
tion (which is much less effective). Explosive manufacturers are important as they
assist users in assuring the safe and effective use of their products for the specific
application.

EXPLOSIVE DENSITY

The density of a material is its mass per unit volume. In the field of explosives,
density is sometimes used in calculating the explosive’s performance in wet con-
ditions. However, the most critical role of density is in determining the explosion
loading of the blast holes. Knowing how many kilograms of explosives fit into each
metre of drilled blast hole allows blasters to create optimum patterns and results.
There exists a “critical density”, which, if exceeded, can cause unwanted results
in the form of poor detonation. This can be caused by loading the explosive too
densely or by a chain reaction of shock waves causing high compaction in adjacent
blast holes.

DETONATION PRESSURE

Detonation pressure is the pressure that occurs ahead of a reaction zone in an explo-
sive. This pressure is calculated by using the material density and detonation velocity.
This pressure in the explosive can pulverize nearby material and radiates outward in
the direction of the shockwaves. The explosive pressure variable is used in determin-
ing the primer needed, as the primer detonation pressure needs to be greater than the
explosive detonation pressure.
290 Underground hard rock mining

EXPLOSIVE ENERGY

Explosive energy is the total energy contained within an explosive and the rate of
the release of that energy in the blasting process. Energy is generally on mass, but
explosive energy is dependent on mass, confinement, rate of release of energy, explo-
sive properties and surrounding rock properties and geology. The energy release rate
(ERR) is therefore the main consideration. This is of particular importance in mining
applications where a low energy release may be required. Consider bench blasting,
where pulverization and fly rock are unwanted effects. The correct balance of released
ERR is something all explosive users must carefully consider.

WATER RESISTANCE

The presence of water is an issue that often causes adverse effects in blasting. Products
such as ANFO that are poured into bore holes containing water can be adversely
affected by lessening detonation velocity, sensitivity and strength. The main solution
for blasting in the presence of water is packaging. By confining an explosive in a water-
proof packaging, blasters can then safely use the material in wet conditions without
unwanted effects. Time is also an important factor with regards to water. Loaded
shots should be set off as soon as possible to prevent explosives from becoming satu-
rated by water. Examples of water resistant explosives are shown in Figure 3.86.

SENSITIVITY

The ease of which an explosive can be detonated is known as its sensitivity. In blasting, the
cap sensitivity will want to be known in order to determine the correct priming method.
However, a large factor of sensitivity comes in the form of safety. Knowing how easy an
explosive can be accidently initiated is of the utmost concern. Friction, impact and shock
are three variables that are tested on an explosive’s sensitivity. These considerations are
taken into account for the storage, transportation and application of explosives.

SENSITIVENESS

Sensitiveness is the measure of an explosive’s ability to propagate outward. This


parameter is important in situations where unplanned propagation would cause prob-
lems. Blasters want adjacent bore holes to shoot independently based on planned
delays and not via propagation. ANFO and ANFO-based water-gels are two common
explosives used that have low sensitiveness.

FUMES

The main gases formed upon detonation are carbon dioxide, nitrogen and water vapor.
These gases are relatively harmless in most cases. Some toxic gases, however, can form
upon detonation including carbon monoxide, ammonia and nitrogen oxides, depend-
ing on the particular explosive used. These are commonly referred to as fumes. In
surface mining, these fumes are quickly dispersed and cause little concern due to the
open environment. However, in underground mining, significant ventilation is needed
to carry these unwanted fumes, along with possible methane and dust, away from the
blasted face to the surface. Fumes are present in virtually all blasting applications and
their quantities are increased due to poor priming, water saturation, mismatching of
charge to blast hole and reactivity to the surrounding rock. Figures 3.87 and 3.88 give
Underground hard rock mining 291

Figure 3.86 W
 ater-resistant explosives. (a) A permitted waterproof explosive (refer to 3.8.1.5);
(b) A waterproof emulsion explosive
Reproduced with kind permission from Orica

Figure 3.87 Carbon monoxide production in relation to fuel oil in ANFO (Rowland & Mainiero, 2000)
292 Underground hard rock mining

Figure 3.88 N
 itrous oxides and nitrogen dioxide in relation to fuel oil in ANFO (Rowland & Mainiero,
2000)

the production of carbon monoxide and nitrous fume at standard temperature and
pressure (STP) as a function of the additional of fuel oil.

FLAMMABILITY

Flammability is the degree at which an explosive can be initiated by heat alone. This is
important for the safe storage and transportation of explosive materials. Materials such
as water gels have a very low flammability compared to materials such as TNT. Testing
is conducted to measure an explosive’s flammability and is printed in the material’s
safety data sheet. Even if a material is rated at an extremely low flammability, precau-
tion should always be taken and unsafe practices should be discouraged and prevented.

3.8.3 Drilling and blasting equipment


The drilling of rock and subsequent charging with explosives or installing support
are still the key to almost all mining operations. There are only a limited number of
exceptions, and these include:

• The use of continuous miners and roadheaders in soft rock (such as coal)
• The use of TBMs or similar machines
• The production of dimension stone using cutting or sawing
• Certain experimental techniques (such as ultra-high-pressure water cutting,
plasma cutting or microwave-assisted breaking)

3.8.3.1 Important requirements for rock drilling


DRILLING ACCURACY

To allow for accurate blasting and removal of ore from a stope, engineers, with assis-
tance from geologists and surveyors, design drilling and blast patterns that radiate from
Underground hard rock mining 293

Figure 3.89 Example of ring drilling (Hassell et al., 2015)

each level access. Depths and directions are precisely designed to maximize ore recov-
ery without dilution from the surrounding rock, so that waste is minimized. Stemming
is important so as to optimize the blast and allow for the effects of the non-parallel
holes. An example of ring drilling in a massive orebody is given in Figure 3.89.

HOLE LENGTH

The hole length drilled is influenced by:


• The ore-body geometry or desired excavation profile
• Production needs or excavation targets
• Available drilling equipment
• Drilling accuracy
The effects of drilling accuracy are given on Figure 3.90.

HOLE DIAMETER

The appropriate blast hole diameter is influenced by the following factors:


• Desired production rate
• Bench height
• Desired fragmentation distribution
• Geological structure
• Rock hardness
294 Underground hard rock mining

Figure 3.90 The causes of drilling accuracy and hole length on blasting (Preston, 1995)

• Explosive type
• Site sensitivity
• Excavation technique

Successful drilling and blasting depend on matching both with each other, and the rock
conditions.

3.8.3.2 Types of rock drills commonly used


PNEUMATIC (JACK HAMMER) ROTARY PERCUSSIVE DRILLS

The jackleg drill is one of the most commonly used drills in the mining industry. Its
use of compressed air to drill into hard rock makes it invaluable for those needing a
compact hand-held relatively powerful and manoeuvrable drill. With extension steel,
it can drill relatively efficiently to 3m, typically 5m maximum, with a diameter just
over 50 mm. The components of a hand-held pneumatic drill are shown in Figure 3.91
and an actual drill being used in Figure 3.92.
ROTARY PERCUSSIVE AIR-BLAST DRILLING (RAB DRILLS)

They are similar to the “jack hammer” but on a larger mechanized scale. The rotary
drilling rig uses a pneumatic reciprocating piston-driven “hammer” to drive a heavy
drill bit into the rock. The rotary drill bit is hollow, solid steel and has ~20-mm-thick
tungsten buttons protruding from the steel matrix. The cuttings are pneumatically
forced out of the rods (in the annulus between the drill rod and the hole) and dis-
charged at the hole entrance. Air or a combination of air and foam (if needed)
remove the cuttings depending on the hole depth. The drill rig is known as a down-
the-hole drill and is commonly used in the VCR mining method.
Underground hard rock mining 295

Figure 3.91 The components of a hand-held “jackhammer” (Vardan & Murthy, 2007)

Figure 3.92 J ackhammer drilling in a hard rock tabular platinum mine (note the extensive temporary
support installed)
Reproduced with kind permission from Impala Platinum Holdings Ltd.
296 Underground hard rock mining

REVERSE CIRCULATION DRILLING (RCD)

The drilling mechanism is a pneumatic reciprocating piston, also known as a


“hammer”, driving a tungsten-steel drill bit. RCD utilizes much larger rigs and
machinery, and and hole depths of up to 500m are routinely achieved. RCD ide-
ally produces dry rock chips, as large air compressors dry the rock out ahead of
the advancing drill bit. RCD is slower and costlier but achieves better penetration
than RAB drilling; it is cheaper than diamond drilling and is thus preferred for
most mineral exploration work unless a core is specifically needed.
DIAMOND DRILLS (CORE DRILLING)

Diamond drilling utilizes an annular diamond-impregnated drill bit attached to the


end of hollow drill rods to cut a cylindrical core of solid rock. The diamonds used are
fine to micro-fine industrial-grade diamonds. They are set within a matrix of varying
hardness, from brass to high-grade steel. Matrix hardness, diamond size and concen-
tration can be varied according to the rock which must be cut. Holes within the bit
allow water to be delivered to the cutting face. This provides three essential functions:
lubrication, cooling and removal of drill cuttings from the hole.
Diamond drilling is much slower than RCD due to the hardness of the ground being
drilled. Drilling lengths of 1,000–2,500m is common, and at these depths, the rock is
mainly hard rock. Diamond drilling rigs need to drill slowly to lengthen the life of
drill bits and rods, which are very expensive, and to ensure good core recovery.
Examples of ring drill rigs are shown in Figures 3.93 and 3.94 (after Sandvik).

3.8.4 Explosives and accessories

3.8.4.1 Main underground explosive types


Types of commonly used explosives include:

• ANFO/ANFEX (ammonium nitrate and fuel oil)


• Nitro-glycerine-based explosive cartridges
• Emulsions
• Bulk explosives
• Permitted explosives (for coal mines)

The main explosive used in mining today is ANFO (Figure 3.95). This powerful and
inexpensive explosive has replaced dynamite due to its stability and economic viabil-
ity. ANFO consists of ammonium nitrate and fuel oil. The ammonium nitrate pro-
vides the oxygen and the fuel oil consumes and burns the borrowed oxygen, creating
gas and pressure. ANFO has a low sensitivity, a wanted characteristic in multiple blast
designs that require staged detonation. ANFO prill is manufactured inside towers
that spray a concentrated ammonium nitrate solution at the top. The solution con-
denses as it falls through a mixture of natural gas and air.
Water gels are used in a wide variety of mining applications. This line of products
was invented in the laboratories of Du Pont through research into safer methods of
blasting. They consist of oxidizing salts, fuels and sensitizers dissolved or dispersed in
a continuous liquid phase.
Underground hard rock mining 297

Figure 3.93 Sandvik DL 321 for ring drilling


Reproduced with kind permission from Sandvik Mining & Rock Solutions

Figure 3.94 Sandvik top hammer drilling rig


Reproduced with kind permission from Sandvik Mining & Rock Solutions
298 Underground hard rock mining

Figure 3.95 ANFO prills - ammonium nitrate and fuel oil


Photo courtesy of Dyno Nobel

Underground coal mining has a tendency to produce methane. This gas is highly flam-
mable when mixed with oxygen in the range of 5–16%. Even with the amount of ventila-
tion required underground to prevent the build-up of methane gas, certain explosives
are used to minimize potential hazards. These types of explosives are referred to as
permissibles. These explosives generally require a flame retardant mixed within the
product to minimize the flame’s duration and/or volume.

3.8.4.2 Initiators and accessories


Explosive accessories are very important to ensure the most cost-effective blasting
result and typically include:

• Fuses
• Igniter cord/trunkline/detonating cords (e.g., Cordex)
• Detonators (initiators):
• Conventional detonator (blasting cap)
• Shock tube detonator/non-electric initiation systems (NONEL) – advantageous
because the system is not affected by stray electrical currents (and lightning)
• EDDs – these have the advantage that the blasting circuit can be tested before
detonation, ensuring improved reliability
• Boosters: Boosters are used to maintain or intensify the explosive reaction at cer-
tain positions along a generally long blast hole

Typical boosters and detonator cord is shown in Figure 3.96.


The range of typical pyrotechnic and electronic detonators are shown in Figure 3.97
(after MSHA, 2021).

3.8.5 Underground blasting


An optimally designed blasting round is one that:
• Minimizes overbreak (unplanned dilution)
• Achieves a maximum advance per blast
• Is the most cost effective overall
Underground hard rock mining 299

Figure 3.96 (a) Components of detonating cord (PETN or pentrite is an explosive); (b) A typical
booster
Booster photograph reproduced with kind permission from Orica

Blasting influences costs significantly because of the following aspects:

• Overbreak and dilution


• Face advance/blast (productivity)
• Size distribution of the blasted rock
• Safety and support costs

Blasting results are influenced by:

• Quality of the drilling (e.g., burden, length, orientation)


• Explosive and accessory performance

Figure 3.97 D
 etonation construction options [Mine Safety & Health Administration (MSHA) Safety
Alert, January 29, 2021]
300 Underground hard rock mining

• Explosive consumption
• Blast hole timing (sequencing)

Controlled blasting techniques are useful to:

• Ensure the accurate shape of the excavation (limit overbreak)


• Reduce the rock to be removed after each blast
• Minimize the support needed
• Minimize blast damage
• Improve safety

Charging the face can be done manually or mechanized and depends mainly on the
dimensions of the excavation. Figure 3.98 shows a highly mechanized charging rig
with a hydraulic platform to reduce the charging time to a minimum.

3.8.5.1 Controlled blasting techniques


Controlled blasting techniques are useful for limiting blast damage around the exca-
vation perimeter, which can reduce the need for extra support. This is particularly
significant at depth. Controlled blasting techniques include:

• Smooth wall blasting – The spacing between the outer perimeter holes is made as
small as practical and de-coupled explosives are used (refer to Figure 3.99).
• Pre-splitting – Its mechanism is to create a free surface around the excavation
perimeter first before the main blast. This tends to reduce the damage caused by
the main blast. This is not a reliable method in higher stress regimes.
• Post-splitting – This is essentially a compromise between smooth wall blasting
and pre-splitting. The blast is fired in the normal way, but the final perimeter holes

Figure 3.98 A mechanized charging rig


Reproduced with kind permission from Orica
Underground hard rock mining 301

Figure 3.99 Example of well controlled smooth wall blasting


Photo courtesy of Dyno Nobel

are blasted simultaneously using light/decoupled explosives. It is difficult to main-


tain the trunkline (connection) between the outer holes during the main blast, and
the post-split blast is often taken later (which is not very productive). EDDs make
this more practical and productive.
• Line drilling – This system involves a single row of closely spaced uncharged
holes (typically 10 cm apart) along the excavation line. This provides a plane of
weakness to which the primary blast can break. It also causes the shock waves
generated by the blast to be partially reflected, which reduces the shattering and
stressing of the rock around the perimeter of the excavation.

3.8.5.2 Tunnel excavation by blasting


The issue in tunnel development is to develop a free face/surface to blast the rock into.
This is achieved by two common methods:

• A burn cut
• A wedge cut
302 Underground hard rock mining

Figure 3.100 Examples of burn cut blast hole patterns (Worsey, 2001)

BURN CUT HOLES

The burn cut consists of two sets of holes: one where the burn holes are heavily charged
with explosives and the other usually larger diameter holes are left un-charged (to cre-
ate some needed voids for the blasted rock to expand into). This effectively creates a
“cylindrical void” that acts as the free surface for the other charged holes. The burn
cut is generally located near the centre of the excavation profile and typical examples
are given in Figure 3.100 (after Worsey, 2001).
The rock in this area is basically pulverized by the explosives and creates a free sur-
face for the remainder of the holes.
WEDGE CUT

The wedge cut holes are drilled to remove a wedge of rock first as shown in Figure 3.101.
Advantages of wedge cut over burn cut:

• Effective for all rock types


• 1 diameter drill steel needed
• Ideal for wide tunnels (maximum advance is half the tunnel width)

Disadvantages:

• Depth of pull (advance rate) is limited


• Flyrock from blasting the wedge can be a problem

A typical blast round therefore consists of:

• Cut holes
• Easer holes
• Breaking holes
• Perimeter holes (sometimes called lifters)

Figure 3.102 shows the typical blast holes drilled and ignition sequences for tunnels.

3.8.5.3 The effect of high stresses on blast fracturing


Blast fractures mainly propagate in the direction of the major principal stress, see
Figure 3.103.
Underground hard rock mining 303

Figure 3.101 Typical wedge cut blast patterns


Source: Dyno Nobel

Figure 3.102 Standard tunnel blast patterns and initiation sequences


304 Underground hard rock mining

Figure 3.103 The effect of high stresses on blast fracture propagation (Worsey, 2001)

In deep mines, this affects the spacing of the lifters and the spacing needs to be
reduced between the holes.
Special blast rounds are needed to create drawpoints or blast rings in mining methods
like longhole open stoping. Typical blast designs are shown in Figures 3.104 and 3.105.

3.8.6 Blast monitoring


Underground blast monitoring is essential for:

• Health and safety first


• Blasting efficiency (underbreak or overbreak)
• Blast vibrations
• Rock fragmentation
• Gas production

This is best undertaken by visual means or instrumentation using say accelerometers


to measure blast vibrations. Most explosive and accessory suppliers also offer services
to help mines optimize their blasting.
Blasting in the vicinity of built-up areas can cause damage and inconvenience to
residents. Frequently, however, some of the damage is historical and has been caused
by natural settlement with time and/or poor building practices or materials.
To avoid damage under most conditions, the peak particle velocity should be kept
below 5 cm/sec. It should be noted that local geology also has a significant role as the
velocity can be changed in a particular direction.

Figure 3.104 Fan drilling to create a cave mine drawpoint (Tucker et al., 2016)
Underground hard rock mining 305

Figure 3.105 Blasting a series of large fan hole rings on a massive mine (Hustrulid & Bullock, 2001)

The assessment of inconvenience is more difficult and subjective. Having open and
regular discussions with potentially affected residences is the most effective method
to reduce complaints.
The effects or air-blasts on surface and underground can be very hazardous and
severe (Figure 3.106). More severe air-blasts can also be formed as a result of, for
example, major underground collapses. Correct stemming has been found to be an
effective way of reducing the air-blast.

3.9 Support design and application

3.9.1 Hardrock support design

3.9.1.1 The need for support and it’s functions


Ortlepp (1989) defines the fundamental requirement of a support as follows: “To keep
the blocks or fragments in position, without itself developing such high internal
stresses as to be destroyed in the process”.

Figure 3.106 The causes and effects of blast vibration (Zhou et al., 2015)
306 Underground hard rock mining

The design of tunnel support must take account of the “human factor”. This means
that the support system must make allowance for the following possibilities:

• The tunnel may not always be adequately “made safe” (i.e., not all the loose rock
may be removed after each blast).
• The development crews may not always be able to identify potentially hazardous
conditions.
• Unplanned overbreak from poor blasting or high stresses may effectively enlarge
the excavation dimensions.
• The support may not always be installed to the planned standard for any number
of reasons, but mainly a lack of training or trying to save time.

The installed support must serve the following functions:

• Prevent local rockfalls


• Reduce the long-term tunnel wall deterioration with time
• Limit unplanned over-break during the tunnel development
• Maintain the overall useful cross-sectional area of the tunnel
• Prevent protection against violent seismically induced movement
• Be cost effective over the planned life of the excavation

All types of support fall into one of two broad categories:

• Passive support by definition is unstressed/tensioned during installation and only


develops a reactive force when rock deformation takes place. The magnitude of
the reactive force depends mainly on:
• The amount of the rock movement
• The stiffness of the support
• The locking or contact mechanism between the support and the rock (in the
case of rock anchors)
• Active support by definition is stressed/tensioned during the installation process.
The support therefore exerts and maintains some restraining force on the rock,
the magnitude of which depends on:
• The stiffness of the support
• The tension imparted during the installation
Support is not needed unless the rock has failed (Figure 3.107). Failure could be caused by:

• Excavation size and shape


• Local geology
• Weak rock and/or high stresses or stress changes
• Blasting
• Weathering/water

Failure in strong rock formations can still occur as shown in Figure 3.108, which
appears to be due to adverse structure in the pillar, rather than high stresses.
Underground hard rock mining 307

Figure 3.107 No support is required in a very large and competent room and pillar limestone mine
Reproduced with kind permission from East Fairfield Coal Company

Figure 3.108 Pillar slabbing due to adverse geological structure of the pillar
Reproduced with kind permission from East Fairfield Coal Company
308 Underground hard rock mining

3.9.1.2 Support design overview


Support of excavations is not a trivial exercise because of the risks associated with
getting it wrong, such as:

• Under-designed – the worst-case scenario as this can result in loss of life in extreme
circumstances.
• Over-designed – this has no additional safety risk but can be a serious economic
risk as too much support has been installed.

The stress tensor obviously is a major factor as are the rock properties. Typically for
shallow excavations, the installed support is designed:

• To hold dead weight


• To be rigid
• To limit and control closure (pillars are useful too)

The challenges in weak rock are also significant and the following should be considered:

• These excavations are difficult to analyze as failure is typically progressive and


weathering can also be a factor.
• Even relatively low stresses can cause failures.
• Typical rock types include mudstones, siltstones, tuffs and some especially lami-
nated shales.
• Deformation monitoring is essential and shotcrete, cast concrete and arches are
commonly used.

For deep excavations, the following is typically needed:

• Support the excavation (failed) skin.


• A yielding support system is needed to handle the relatively high rock deformation.
• If seismicity is possible, then the support must be able to absorb energy.
• Try to avoid leaving pillars as they act as stress concentrators, which is generally
undesirable.

Keyblock theory (Shi, 1988) is useful for considering the support needs in hard rock
mines. A keyblock is any discrete block of rock, bounded by geological discontinui-
ties, blast or stress-induced fractures that could fail under gravity alone, if not sup-
ported. An example is shown in Figure 3.109.
Beam building is another common hard rock (and especially soft rock) support
need usually in sedimentary rocks with prominent parallel bedding planes as shown
in Figure 3.110.
The angle that a rockbolt intersects a failure plane is important as shown in Figure 3.111
(Stillborg, 1986). An angle close to 90° to the failure plane is the most efficient.
Hoek introduced the concept of rockbolts creating a reinforced beam of thick-
ness “t” as shown in Figure 3.112. To create this and hence have a support system
(as opposed to just random rock anchors), the spacing “s” must be less than the
rockbolt length “L”, and the closer the spacing for a given rockbolt length, the
Underground hard rock mining 309

Figure 3.109 An example of hard rock keyblock support

Figure 3.110 Other examples of hard rock beam building

Figure 3.111 The importance of bolt inclination to a shear plane (bolt B is more effective)
310 Underground hard rock mining

Figure 3.112 An artificially created rock reinforced beam using rock anchors (Hoek & Wood, 1987)

thicker the reinforced arch produced. A full pressure arch around a unnel is shown
on Figure 3.113 (Hoek et al, 1980 and 1987).
Figure 3.114 shows the different rock conditions typically found underground and
needing support in mainly hard rock applications.
Always consider the temporary support needed during the excavation process
before the final/permanent support is installed. This support needs to be removed
after the permanent support has been installed and should be done from a position of
safety (possibly under the protection of the newly installed permanent support).
For a complete and appropriate rockbolting system design, the following parame-
ters must be determined properly:

• Bolt type
• Bolt length
• Pattern and spacing bolts
• Bolt diameter and anchor capacity
• Whether pretension should be applied or not. If pretension is desirable, what is the
appropriate magnitude of the pretension?

Figure 3.113 A pressure arch example (Hoek & Wood, 1987; Hoek & Brown, 1980)
Underground hard rock mining 311

Figure 3.114 Underground rock conditions around excavations (Stillborg, 1986)

3.9.1.3 Design approaches


The basic considerations are:

• No support is required if the stresses do not exceed the strength of the rockmassm,
and there is no blast induced damage.
• The rockmass is weakened by geological features (faults, joints, bedding planes, etc.).
• Support needs increase with increased depth and more significantly with excava-
tion span.

Field stress and future stress changes due to mining are critical:

• Vertical stress (σv) depends on the overburden depth (h in km), the gravity con-
stant (g) and rock density (ρ):

σv = ρ * g * h in MPa

• The magnitude of the horizontal stress, however, is difficult to estimate especially


at shallow depths where it may exceed the vertical stress by a factor of ×3. At
depth, it tends to approach ½ the vertical stress.

The following (Figure 3.115) is a useful design procedure to follow after Bieniawski
(1984).
The feedback loop is an absolutely critical activity, and the design must be checked
regularly to ensure that it remains safe, cost effective and meets the actual ground
conditions encountered underground (in situ).
312 Underground hard rock mining

Figure 3.115 The overall design methodology (Bieniawski, 1984)

The choice of tunnel support can be based on various methods and principles:

• It is reasonable to assume that if a certain support system has been successful


under similar conditions, it will work again.
• If there is a chance of major seismicity in the vicinity of the tunnel, a yielding sup-
port system must be installed with a suitably designed fabric between the anchors
(such as meshing and lacing).
• Economics must always be a major consideration in any tunnel-support design.
• Computer analysis packages (both elastic and inelastic) can be utilized, especially
if some form of calibration exercise has been conducted, based on underground
measurements, observations and monitoring.
• Keyblock theory or any applicable system of rock-mass characterization can be used.
The Rockwall Condition Factor (RCF) after Wiseman (1979) can also be useful.

RCF = (3σ1 − σ3 ) / Fσc

Where σ1 and σ3 are the major and minor principal stress components acting normal
to the tunnel, F is an empirical rock-mass condition factor and σc is the uniaxial com-
pressive strength of the country rock.
In highly competent rock, F = 1, and this value should be decreased for incompetent
rock to about 0.5. The size of the tunnel also influences the value of F and, if a large
tunnel cross-sectional area is planed (say more than 36m²), F should be reduced by at
least a further 10%.
If the RCF is less than 0.7, the underground stability will be good, but if the RCF is
greater than 1, the underground stability will be very poor.
A typical tunnel design process would be as follows (Figure 3.116).
Underground hard rock mining 313

Figure 3.116 A tunnel design process

3.9.1.4 A successful support design


A successful design:

• The measure of success is not only a safe solution, although this is key, the solu-
tion adopted must also meet the functional requirements of the excavation and
also represent a cost-effective solution.
• Engineering judgment based on available information is critical, and the design
must be revisited and modified as more information becomes available.

3.9.2 Rockbolts and cable bolts


The following are some of the common rock anchors used in hard rock mines.

3.9.2.1 Mechanical expansion shell


This type of rockbolt is:

• An active point anchored bolt


• Can be post grouted to become an effective permanent support
• Loses tension mainly due to adjacent blasting, and bolts should be re-tensioned
after a nearby blast
• Frequently used for spot bolting or as a temporary support
• Should not be used in soft or friable ground

Figure 3.117 shows an example of a mechanical (ungrouted) rockbolt.


Two types of mechanical anchors are used: a bail type shell and a standard (sleeve)
type shell, as shown in Figure 3.118.
A mechanical rockbolt installed at 30° off the perpendicular may provide only 25%
of the tension produced by a bolt equally torqued that is perpendicular to the rock
face, unless a spherical washer is employed.
314 Underground hard rock mining

Figure 3.117 A mechanically anchored rockbolt (modified from Stillborg, 1986)

3.9.2.2 Deformed bar/rebar/ripple bar


This is the most common type of rockbolt used (Figure 3.119). It is commonly called
rebar (from REinforcing BAR).
It can be:

• A passive support (anchored with a single-speed cementitious or resin grout)


• An active support (anchored with a mechanical shell with cement or resin grout or
by two-speed resin). The rockbolt is then tensioned during installation.

Figure 3.118 T
 he two main mechanical shell types (with a third modified option) (University of
Wollongong, 2020)
Underground hard rock mining 315

Figure 3.119 Typical grouted rebar (modified from Stillborg, 1986)

The cement grout can be pumped in or capsules/cartridges can be pre-inserted, or


the resin is usually inserted in the form or capsules/cartridges. The grout quality and
correctly (completely) filling the hole with grout are the most important considera-
tions. Grouted anchors are used as a permanent support. Using grouts in the presence
of running-water can be a challenge and it may be better to consider alternative sup-
ports (or even water sealing if extreme).
Holes drilled for cement-grouted bolts should be 12–25 mm (½–1 inch) larger in
diameter than the bolt. The larger gap is especially desired in weak ground to increase
the bonding area.

3.9.2.3 Friction bolts (split sets)


Friction bolts (Figure 3.120) are:

• A passive support as load reaction depends on the friction of the spring steel
against the rock periphery (friction rock stabiliser)
• Rapid and easy to install
• Prone to corrosion, and if such conditions exist, stainless steel units, or post
grouting should be used.

The hole diameter is important, and a simple “go, no-go” device for the drill steel
should be used.
For each 30 cm (1 foot) of friction bolt (split-set) installed in the rock, there is about
1 tonne of anchorage support.
316 Underground hard rock mining

Figure 3.120 A friction rockbolt (modified from Stillborg, 1986)

3.9.2.4 Swellex bolts (Atlas Copco)


The following are relevant for this type of rockbolt (Figure 3.121):

• Quick installation
• Requires a high-pressure pump
• Semi-active support
• Can corrode (stainless unit available at a cost premium)
• Mainly used for high-speed mechanized development
• Load reaction depends mainly on the pump pressure (that is used to “inflate” the
deformed tube after it has been placed in the hole)
• Rapid and easy to install, especially in mechanized development
• Can be made to “yield”/deform

Figure 3.122 shows how a Swellex bolt is fitted in the drill hole then inflated using a
very high-pressure water pump.

3.9.2.5 Resin-grouted bolts


A typical active fully resin grouted rebar is shown in Figure 3.123.
The resin cartridge has two compartments (Figure 3.124): one has the filled polyes-
ter resin (mastic) and the other has the filled catalyst/activator. When the components
are mixed by spinning the rockbolt through them, they mix together and set, with
typically 30 to 50 spinning revolutions. The resin typically shrinks onto the rockbolt
surface and creates an effective mechanical key with the rock around the hole. The set
resin mix does not really adhere to the rock.
Underground hard rock mining 317

Figure 3.121 A Swellex bolt (modified from Stillborg, 1986)

Installation practices for passive and active bolts mainly in coal mines are summa-
rized below:

• The installation of traditional fully grouted passive rebar rockbolt is relatively


straightforward and simple. On installation, a single resin cartridge is inserted
into the drill hole. As the rockbolt is inserted into the drill hole, the resin cartridge
bursts, and is thereafter spun for an allotted number of seconds as recommended

Figure 3.122 S ection through an uninflated then an inflated Swellex in a pipe. (a) Uninflated Swellex
bolt; (b) Inflated Swellex bolt
Photograph taken by Spearing using samples kindly given by Atlas Copco.
318 Underground hard rock mining

Figure 3.123 U
 sing two setting speeds of resin to make an active (tensioned) rockbolt (modified
from Stillborg, 1986)

by the manufacturer and held via the up-thrust of the bolter unto the bolt. During
this time, the resin sets and loading to the bolt is attributed to subsequent rock
movement overtime.
• For two-speed active torque tension rockbolts, two cartridges are inserted into
the drill hole: a small and fast-setting resin cartridge followed by a larger and

Figure 3.124 T
 ypical resin cartridge components (the black is the resin/mastic component and the
white is the catalyst component, they are separated into two components)
Photograph taken by Spearing
Underground hard rock mining 319

slow-setting resin cartridge. During the installation, the fast-setting resin gels first
and the bolt is tensioned using a threaded bolt and nut; the tension is kept in the
bolt as the slower resin sets. A new development is a single cartridge containing
two resin gel times (fast and slow). This also reduces the exposure of the bolter
operators to the unsupported top, as only one longer resin cartridge is needed
instead of two for a tensioned rebar bolt. This also improves bolting productivity
significantly as the time to insert the cartridges is obviously reduced.
• For use of a partially grouted resin-assisted mechanical anchor rockbolt, a
mechanical shell is attached to the end of a bolt with the use of a partially grouted
resin column (generally a partial column due to the high insertion back pressures
caused by the shell). The bolt is tensioned to fix the shell, and impart a tension,
and then the resin sets, thus locking the tension in place.

Important operational parameters for the successful and cost-effective use of res-
in-grouted rockbolts are:

• Standard practice is to drill the bolt hole an inch longer than needed.
• The annulus (annular gap) between the rockbolt and the hole is defined as half the
diametric difference between the drill hole diameter and the rockbolts diameter.
Keeping this low can facilitate the mixing of the resin components during instal-
lation. The optimum annulus is 3–4 mm and more than 8 mm should be avoided.
• As mentioned, manufacturers typically recommend 30–50 rotations during bolt
rotation, which with typical bolter speeds equates to 3–5 seconds for fast-setting
cartridges. Fast cartridges are used so installation time per unit is minimized, as
in room and pillar mining, support is the production bottleneck.
• After spinning the bolt, if it is passive, it is held for several seconds then the bolter
boom is lowered; if tensioned with the mechanical shell, it is tensioned (torqued)
immediately; if two-speed resin, it is held for typically 3–5 seconds and then
torqued by the rockbolter.

As stated, holes drilled for resin bolts should be about 6–8 mm larger in diameter than
the bolt. If it is increased to around 10 mm, the pull-out load is not significantly affected
(the spin time may need to be increased though) but the stiffness of the bolt/resin assem-
bly is lowered by more than 80%, besides wasting money on unnecessary resin.
A relatively new resin cartridge development to achieve a pre-stressed rockbolt is the
use of a two speed resin cartridge where a single cartridge contains resin mixes with
two setting times, usually a fast and a slow gel time. This is faster to install and more
convenient, but the cartridge must be installed correctly with the fast set portion into the
hole first. The different set speeds are achieved by using a stronger or a weaker catalyst
component, and the fast part of the cartridge is coloured differently to help identify it.

3.9.2.6 Cable bolts


Long cable bolts (Figure 3.125) can be more easily used than solid bars, and this is
highly beneficial where space is limited. They have a high tensile load-bearing capacity
and can be made to any length. They can also sustain more deformation that standard
rebar rockbolts. They can be active and passive and can be manufactured to withstand
very high loads (up to 100 tonnes each).
320 Underground hard rock mining

Figure 3.125 An example of the many cablebolt types commercially available


Reproduced with kind permission of Jennmar Corp.

3.9.3 Specialized support


The following are examples of rock anchors needed for more specialized applications
such as under seismic conditions, weak ground or anchors used to achieve a specific
result (such as much higher productivity).

3.9.3.1 Yielding rockbolts


In areas where seismicity or large quasi-static deformation is expected, yielding rock-
bolts can and should be used. The main design consideration is not tensile capacity
but energy absorption. There are numerous different designs of rockbolts used under
rockburst conditions, and only a few examples are discussed.
THE CONE BOLT (SOURCE: CHAMBER OF MINES RESEARCH ORGANIZATION, SOUTH AFRICA)

The Cone Bolt (Jager, 1992) was one of the first successful rapid yield rockbolt for
seismic conditions (Figure 3.126). Figure 3.127 shows a Cone Bolt being tensile tested
under slow and rapid (3m/s) conditions.
GARFORD DYNAMIC SOLID BOLT (HEDRICK, 2010)

It is a solid bolt that has a mechanism that allows it to yield in a stable manner and
thus absorb seismic rapid movement (Figure 3.128). The solid bolt can be installed in
both resin and cement grouts. The polyethylene sleeve acts as a debonding agent that

Figure 3.126 The smooth de-bonded Cone Bolt (Jager, 1992)


Photo from COMRO in South Africa which no longer exists
Underground hard rock mining 321

Figure 3.127 The Cone Bolt under large, slow and rapid deformation

Figure 3.128 The Garford Dynamic Solid Bolt, yield mechanism and performance
Reproduced with kind permission from Garock, Australia
322 Underground hard rock mining

Figure 3.129 The D Bolt


Reproduced with kind permission from Normet

allows the solid bolt to slip through the sleeve, which acts as the dynamic device. It is
mechanical, therefore enabling repeatability in regard to the energy absorption pro-
cess. The dynamic device is factory installed so good quality control is assured.
THE D BOLT (NORMET, 2006)

The D Bolt (Figure 3.129) is a rockbolt comprised of a smooth steel bar with a number
of deformed sections that act as specific anchor points along the bolt’s length. The
collar end of the D Bolt is threaded and is designed to be used with standard fittings.

3.9.3.2 Self-drilling anchors


Self-drilling anchors are often used in weak or fractured ground where if the drilling
rods are removed, the hole may partially or totally collapse, making rockbolt insertion
impossible. The self-drilling anchor basically has a single use drill bit and the drill rod
acts as the rockbolt member after drilling too. Figures 3.130 and 3.131 show examples
of short and long self drilling rockbolts.

Figure 3.130 Self-drilling friction bolt


Reproduced with kind permission from Jennmar Corp.

Figure 3.131 Self-drilling grouted rockbolt


Reproduced with kind permission from Jennmar Corp.
Underground hard rock mining 323

Figure 3.132 The concept of spiling (Forbes et al., 2018)

3.9.3.3 Spiling bolts


Spiling bolt (spiles or forepoles) provide a protective canopy, ahead of the excavation
advance where the rock mass is weak or very friable. Installing these is costly and time
consuming so it is only used when the ground conditions dictate the need. A layout is
shown in Figure 3.132.

3.9.4 Shotcrete

3.9.4.1 Shotcrete history


Shotcrete (or sprayed concrete as it is often called) is very commonly used in tun-
nelling projects and on metal and non-metal mines as part of a support system. The
development history is as follows:

1907 – Dr. Carl Akeley, the well-known US naturalist, first developed a machine
to spray concrete (Figure 3.133). He needed to develop a machine to spray the
concrete over his wire-framed dinosaurs that he was creating for a park.
1910 – First introduced to the public at the Cement Show, Madison Square Gardens,
New York.
1915 – The concept is improved by the Cement Gun Company, later to become
Allentown Gun Company. Karl Akeley also registered the term “gunite” for
a sprayed mortar mix.
1920s – Shotcrete was used in the US to fireproof underground drifts. First recorded
underground application was in the Mersey Tunnel in 1928 (Figure 3.134).
1930s – The term shotcrete was introduced by the American Railway Engineering
Association.
1940s – Coarse aggregates were introduced into the mixes.
1954 – Anton Brunner, an Austrian engineer, replaced heavy steel and tim-
ber support with shotcrete in a tunnel at the Runserau Power Plant. This
was probably the first application that was later named the New Austrian
Tunnelling Method (NATM) in the 1960s.
1955 – The wet shotcrete application process was developed.
324 Underground hard rock mining

Figure 3.133 The first shotcrete in 1910 (photo taken by the late K. Akeley) (Sauer & Gold, 1989)

Figure 3.134 F irst use of shotcrete underground in the Mersey tunnel, Liverpool, UK (opened in
1934). Full reference unknown.
Underground hard rock mining 325

3.9.4.2 The dry shotcrete machine and process


American Concrete Institute (ACI) defines dry mix shotcrete as: “The shotcrete in
which most of the mixing water is added at the nozzle”.
The dry mix process involves the mixing of cement and aggregates in the designed
proportions before supplying it to the shotcreting equipment. The final mixed ingre-
dients are then placed in the feed hopper. During the shotcreting operation, the
mix is taken from the hopper to the nozzle through the delivery hose by using com-
pressed air.
Once the dry mix reaches the application nozzle, water under high pressure is
sprayed to the mix through a perforated ring attached to the nozzle assembly. During
spraying therefore, the water thoroughly “wets” the mix, making it a concrete paste
that is ejected at high velocity against the surface to be shotcreted.
The amount of water added for the dry mix shotcrete process is controlled by the
nozzleman. This is the key operation and relies on the experience and training of
the nozzleman. The water is controlled by means of a valve situated in the nozzle
assembly.
The dry mix shotcrete is typically applied in areas where the total requirement is rel-
atively low and access is difficult. Figure 3.135 shows a section through a dry shotcrete
machine and Figure 3.136 shows an actual dry shotcrete machine.
The rotor is similar in design to that of a revolver, dry material enters a chamber,
the multi-chamber mechanism continuously rotates and the dry material is discharged
into the shotcrete hose using compressed air at 180° from where it is filled with the dry
mix (Figure 3.137).

Figure 3.135 Section through a typical dry shotcrete machine


326 Underground hard rock mining

Figure 3.136 A dry shotcrete machine


Reproduced with kind permission from Normet

The maintenance of pump and ancillaries is critical. The following are the most
important aspects that need to be checked:

• Keep the hopper free of concrete after use


• Check the re-mixer is working and in the right direction
• Wear plate and ring must be tight with no gap
• Swing tube action must be free

Figure 3.137 The rotor, filler plate and discharge wear plate.
Reproduced with kind permission from Normet
Underground hard rock mining 327

• No residual concrete build-up in lines and hoses


• In dry equipment, keep the seals, etc. in good order

Dry shotcrete machines tend to be very compact and with relatively low output.

3.9.4.3 The wet shotcrete process and pump


The wet mix process involves mixing all ingredients to form a complete concrete mix
(including the correct water). The shotcrete mix is therefore prepared before placing it
in the shotcreting hopper. The delivery pump is usually either a positive displacement
or worm pump.
The process involves forcing the wet mix to the nozzle through a high-pressure
delivery hose using compressed air. The mix is then again shot at high velocity onto
the surface to be shotcreted. A worm pump or a piston pump are generally used for
wet shotcreting (Figure 3.138).
Wet shotcrete pumps have much higher capacity, and many different variations
exist from small sled or trailer mounted units (see Figure 3.139) with capacities around
6–8m³ per hour to large carrier-mounted units with capacities in excess of 24m³ per
hour. Figure 3.140 shows a large wet shotcreter on surface and Figure 3.141 shows
a machine placing shotcrete underground safely, as the operator is remote from the
actual placement.
The wet pump can also be retrofitted onto any suitable mobile carrier.
Special “fit for purpose” application arms can be made for wet shotcreting. A nozzle
can be fitted for example around a circular beam in a sinking shaft to readily apply
shotcrete to the newly excavated shaft. This can be done as the final lining or as a tem-
porary support before pouring the cast concrete rings.

Figure 3.138 The common two wet shotcrete pumps (Vandewalle, 1998)
328 Underground hard rock mining

Figure 3.139 A trailer-mounted wet shotcrete machine

Figure 3.140 A wet shotcrete machine with a boom on a carrier


Reproduced with kind permission of Normet

Figure 3.141 P
 lacing high volume and quality wet shotcrete underground (note the absence of dust
and low rebound)
Reproduced with kind permission from Normet
Underground hard rock mining 329

Figure 3.142 Section through a peristaltic pump

Accelerators are routinely added at the discharge nozzle of the shotcrete and the
dosage must be accurate and consistent. A peristaltic (hose or Bredel) pump is often
used for this application (Figure 3.142).

3.9.4.4 Selecting the optimum shotcrete process for the specific application
The selection of the most appropriate process depends on local conditions, placement
rates required and logistical considerations. Table 3.12 (Melbye, 2006) gives a compar-
ison between the methods.
It should be noted that if fibres (especially steel fibres) are used in the shotcrete mix, then
the dry process is to be avoided, if possible, due to the very high losses of fibres in the rebound.

3.9.4.5 Admixture use


Admixtures are commonly used especially when wet shotcreting for the following reasons:

• To accelerate the rate of strength gain


• To improve the flowability (without adding water which causes more slump and a
lower strength)

Table 3.12 Comparison between the shotcrete processes

Wet process Dry process

Little dust Considerable dust


Low maintenance High maintenance
High capital Low capital
Low rebound (typically about 10%) High rebound (usually more than 25–40%)
Moderate to high placement rate Low to moderate placement rate
(between 3 and 20m³/hr) (up to about 5m³/hr)
Low transport distance (up to about 200m) High transport distance
Moderate to high placed quality Moderate to low placed quality
330 Underground hard rock mining

• To increase the single pass thickness


• Delay the initial hydration for easier transportation to site

Don’t use an admixture without testing first and ensure that the dose rate is correct
during spraying: this means a reliable and accurate dosing system and pump.
In fact with shotcrete, everything associated with the mix and the components
should be first tested and an effective quality assurance system is essential. Cement
is a weak alkali so checking that the aggregate is not acid-producing is important.
Quality control tests should be:

• Reliable
• Meaningful
• Timeous (quick for most tests)
• Simple (so the operators can do them as part of their jobs)
• Relatively inexpensive

The main parameters checked in any shotcrete quality programme should be asso-
ciated with design compliance (mix design, water content, bond and strength) and
sprayed design thickness. It is, however, totally unacceptable to have a quality control
system in place, but fail to take adequate actions, if non-compliance is identified.

3.9.4.6 Shotcrete nozzles


Shotcrete nozzles are important to ensure good water mixing in dry shotcreting,
accelerator addition and past and good compaction of the shotcrete being sprayed.
A nozzle assembly is shown in Figure 3.143.
The nozzle assembly must have:

• An adequate air hose size


• An accelerator hose
• A proper mixing apparatus especially for dry shotcrete to ensure thorough “wet-
ting” of the concrete
• A well-designed nozzle

Figure 3.143 A typical wet shotcrete nozzle assembly


Underground hard rock mining 331

For this to work, the compressor has to be big enough, meaning minimum 10–12m3/
minute of compressed air. The pressure is obviously important too.

3.9.4.7 Shotcrete production (in batch plants)


Much is available concerning the design of shotcrete, but simple aspects are frequently
overlooked initially, and these can create major losses, costly delays and final sprayed
linings that do not meet the desired performance.
Shotcrete design must include more than creating a laboratory mix that meets the
strength gain requirements, with locally available raw materials (such as cement, sand,
stone and water) in adequate supply. Whilst this is important, three other equally vital
aspects must not be ignored:

• The fact that the strength must be achieved on the rock (not in the lab)
• The time available for spraying a given volume
• The cost of the mix

Shotcrete is basically concrete so consists of the same components, and when pro-
duced, the same quality control procedures must be applied. It therefore consists of
the following materials:

• Aggregates
• Sand
• Cement
• Water
• Admixtures
• Additives (such as micro-silica and fibres)

The water quality is important and should be free from oils and not heavily acidic (as
cements are alkaline). As a rule, if the water is potable (safe to drink), it is good for
concrete (shotcrete).
A small plant can be set up on the mine site, depending on the application to feed
directly into the shotcrete applicator.

3.9.4.8 Shotcrete transportation from plant to application site


The shotcrete must be transported from where it is produced (batched) to where it will
be sprayed (used).
The two main options are:

• Via a concrete transporter (vehicle)


• Through a pipeline

The selection is mainly dependent on the mine site, the volumes of shotcrete (concrete)
needed, the shotcrete locations and the ramp and tunnel access dimensions. If a sur-
face decline is not available, the shotcrete is generally transported to a central location
via a pipeline in a shaft or borehole, then distributed by truck to the specific applica-
tion site. Depending on the tunnel dimensions and shotcrete volume requirements at
332 Underground hard rock mining

Figure 3.144 A low-profile vehicle designed to carry shotcrete (or concrete) underground
Reproduced with kind permission of Normet

any specific time, small or large shotcrete carriers can be used as shown in Figures 3.144
and 3.145.
A pressure reducer (or kettle) is needed when “dropping” concrete (shotcrete) ver-
tically down a pipeline in a shaft or a borehole. Figure 3.146 shows three examples of
effective pressure reducers that can be used.

3.9.4.9 Shotcrete costs


Shotcrete placed costs are seldom if ever accurately estimated. The same is evident for
most underground support systems.

Figure 3.145 A large capacity vehicle to deliver shotcrete (or concrete) underground
Reproduced with kind permission of Normet
Underground hard rock mining 333

Figure 3.146 E xamples of effective pressure reducers for use when dropping concrete or shotcrete
vertically down a pipeline (Spearing, 2002)

The following are some of the more significant cost elements:

• Material Costs: The material cost is generally relatively easy to establish and is
often higher for wet mixes.
• Equipment Costs: The capital cost for the shotcrete and ancillary equipment is
also easy to determine, and it varies significantly depending on the capacity and
on the application process. Wet machines are always significantly more costly
than dry machines.
• Maintenance Costs: The operational cost of the equipment is generally over-
looked but can be substantial. Dry equipment maintenance per m³ sprayed is
between two and four times more than with wet equipment. In Canada and the
US, maintenance (and replacement) costs for the dry machine, hoses and nozzle
are typically around $25/m³ sprayed.
• Material and Equipment Transportation Costs: This important cost is generally
ignored or avoided but can be very significant and should be considered. Mines
that bother to investigate this aspect are generally able to justify the installation of
a shaft pipeline with pumps or “agitator-cars” for the horizontal transportation.
This assumes that the monthly volumes needed are regular and relatively large.
• Labour Costs: The true labour cost involved in the entire process needs to be
considered. When comparing shotcrete against other supports, any rehabilitation
costs should also be estimated.
• Application Efficiency Costs: The rebound must be considered when shotcreting,
and it is not too difficult to estimate. Quality wet shotcreting should be 10% or
less, and quality dry shotcreting should be 25% or more. This must be accounted
for in any shotcrete costing exercise.
• Time-related Costs: These need to be considered if the mining cycle is critical for
a specific excavation development. The effects of lost blasts, due to the support
installation taking too long, are very significant in many cases.
334 Underground hard rock mining

3.9.5 Backfill

3.9.5.1 Introduction to backfilling


Backfilling is a “green mining” technology as it reduces waste on surface and can
reduce surface subsidence and ground water damage.
Backfilling is the process whereby waste material (generally metallurgical tail-
ings or prepared aggregate) is placed back underground into open mine workings
(stopes or panels). It is becoming an essential part of many underground mining
operations. It can also resolve some of the environmental surface impacts, but in
most metal mines, too much metallurgical waste is produced so not all can be placed
back underground.
Backfill is typically used for the following mining applications:

• As a working platform
• As an artificial roof
• As a stope/void filler
• As an artificial sidewall and a pillar replacement
• To reduce surface subsidence

This has many potential advantages for typical underground mining applications,
such as:

• Reduced tailings on surface


• Reduced surface subsidence
• Increased underground extraction
• Reduced rockburst (seismic) damage
• Improved support for mining excavations and/or working surfaces
• Increased worker safety

Typical sources of backfill material include:

• Waste metallurgical tailings that can be used as total or classified material


• Underground waste rock
• Sand or quarried rock
• Industrial waste (such as fly ash)

There are therefore several different basic types of backfill used:

• Rock fill (RF)


• Cemented rock fill (CRF)
• Uncemented classified (hydraulic slurry) tailings fill (CT)
• Cemented classified (hydraulic slurry) tailings fill (CCT)
• High density/paste fill
• Solid backfill (developed by the School of Mines at China University of Mining &
Technology (CUMT)) specifically for longwall mining above sensitive surface areas

The performance of some different backfill types is shown on Figure 3.147.


Underground hard rock mining 335

Figure 3.147 Performance behaviour of different backfill materials (Spearing, 1989)

3.9.5.2 Mining methods using backfill


Backfilling is an integral part of many mining methods, such as:
• Longwall mining (coal and hard rock)
• Cut and fill and undercut and fill
• Longhole open stoping and VCR
• Drift and fill
• Room and pillar, with pillar recovery
LONGWALL MINING

Longwall mining is commonly used for narrow tabular mining at orebody (reef) dips of
usually less than 30°. It can be used in hardrock applications (such as gold mining on the
Witwatersrand Basin in South Africa) or in softrock applications (such as coal mines).
ROOM AND PILLAR MINING

The key to this mining method is the room width (the span between the pillars). This is
generally selected using geotechnical evaluations (including rock characterization meth-
ods) so that the room can be safely mined with the minimum of installed support to cater
for minor features such as joints. The relationship between increasing the room width and
hence increasing the installed support is generally dictated by economics. Tributary area
theory is usually adopted to design the pillars and tends to give conservative results.
Once the primary mining has been completed (i.e., the rooms removed) some form
of total or partial secondary extraction (usually on a retreat basis) can be undertaken
(with or without backfill depending mainly on any subsidence limitations). Total
extraction is seldom achieved as the secondary pillar extraction is usually difficult
and inefficient (hence more costly).
336 Underground hard rock mining

This method of mining is generally applicable in orebodies where the dip is 25° or
less and the orebody and the hangingwall (roof) are relatively competent. Increasing
the pillar dimensions and/or reducing the room width can compensate for poor ground
conditions but overall extraction is significantly reduced.
There are three basic types (or variations) of this mining method:

• Horizontal or flat room and pillar mining


• Inclined room and pillar mining
• Step room and pillar mining

CUT AND FILL/DRIFT AND FILL MINING

In the common form, cut and fill mining involves taking horizontal slices of ore, back-
filling the void created and moving progressively upwards. The method can also be
used to mine downwards, called undercut and fill, and this requires a competent back-
fill that is very costly.
The backfill is used as a working platform and a support for the orebody walls. In
wider orebodies, drift and fill may be used, where horizontal rooms are developed and
backfilled, before the next vertical slice is removed.
Productivity is dependent on achieving the whole mining cycle of drilling and blast-
ing, making safe and supporting, removing the ore and backfilling. The method is
relatively labour intensive and tends to be utilized in the extraction of more valuable
and higher grade minerals (such as gold).
It is generally applied in the following circumstances:

• Steep dip
• Relatively thin orebody widths
• Competent orebody
• Country/host rock can be relatively incompetent
• Can be used at any depth

SUB-LEVEL/OPEN/BLASTHOLE STOPING (AND VCR)

This mining method is used in orebodies with the following characteristics:

• A steep dip (more than the angle of repose of the material)


• Competent footwall (floor) and hangingwall (roof)
• Competent ore
• Regular orebody boundaries

In sub-level stoping, after the ore is extracted, the stope is left empty. Stopes are typ-
ically very large with the largest dimension in the vertical. Pillars are left between
stopes depending on the geometry of the orebody.
Mining in sub-level stoping is carried out from horizontal levels and sub-levels at
pre-determined vertical intervals. The ore is broken by drilling and blasting from the
sub-level drifts and collection from the main level below.
Drilling in sub-level stoping can be carried out well in advance of mining and the
blast hole length drilled is seldom more than 25m. An outline of the method was given
in Figure 3.52.
Underground hard rock mining 337

SUB-LEVEL CAVING

Most caving methods function under the principle that the orebody and the country
rock fracture under more or less the same conditions.
In sub-level caving, the ore is divided into sub-levels with a comparatively small
inter-level distance (usually between 8 and 15m). Sub-level caving is used in steeply
dipping orebodies and deposits with large vertical dimensions. Sub-level drifts must
be basically self-supporting, requiring only spot bolting. The rock in the hangingwall
(roof) must cave and follow the orebody extraction in a continuous cave right up to
surface.
The application of sub-level caving tends to be for the following deposits:

• Where the dilution from the waste can be easily separated from the ore minerals
(for example, by magnetic separation)
• Where the boundaries of the orebody are vague, and the surrounding country
rock is mineralized to some extent

Each sub-level is developed with a regular network of drifts across the entire orebody
section. The ore immediately above each sub-level drift is drilled using a longhole pat-
tern. This drilling is undertaken well in advance of the blasting. The blasting of each
level takes place from the hangingwall on a retreat basis. The ore extraction usually
takes place along an approximately straight front, and this permits the simultaneous
operation of adjacent drifts. Explosive consumption is high because the ore is blasted
against the caved waste.
As broken ore is extracted at the drawpoint, fragmented ore and the enclosing caved
waste displace to fill the temporary void. The success of the draw and hence of the
method itself is determined by the relative mobilities of the fragmented ore and the
caved waste.

3.9.5.3 Backfill transportation and design


There are four main transportation methods (or combinations) used to move the fill
from the point of preparation (generally surface) to the underground workings:

• Hydraulic via pipelines, with the fill as a slurry (low density) or paste (high den-
sity). This is the most common method, but as the slurry density increases towards
a paste, the pressure gradient increase and pipeline tests are advisable before the
diameter is specified. The pressure loss (due essentially to friction) relationship to
transport velocity in a pipe is shown in Figure 3.148.
• Mechanical via conveyor belts or trucks. Trucking and conveyors are commonly
used to transport rock fill. Each has merits of its own and should be evaluating
individually before a decision is made. Considerations that should be made include
the number of areas requiring simultaneous placement, fill tonnage requirements,
access and ventilation considerations.
• Pneumatic via pipelines as a dry (or damped) material. This method of transport
has become unpopular mainly because of the high energy consumption and cost
of using air, the severe wear rate due to impact and the dust (even if the material
is pre-damped).
338 Underground hard rock mining

Figure 3.148 Pressure loss versus velocity for a homogeneous slurry

• Gravity via raises. This method is commonly used to transport rock fill vertically,
and the raises can be lined or unlined. Depending on the strength of the rock, lining
the raise with an impact and wear-resistant liner is often cost effective if the cost
of rehabilitation of an unlined raise is considered. Generally, it is advisable to run
rockfill raises/passes full to reduce wear and try to limit degradation that creates
more fines. This does have the potential to increase the risks of hang-ups in the raise.

An example of a paste backfill plant is given in Figure 3.149.

Figure 3.149 A paste backfill plant example (Belem & Benzaazoua, 2008)
Underground hard rock mining 339

3.9.5.4 Backfill criteria


The following are some of the more common and important criteria and requirements:

• The fill should be placed at the lowest possible cost.


• The risk of fill failure (e.g., liquefaction) must be minimized.
• Early strength development needs to be adequate and meet the design.
• Long-term strength should be sustainable.
• Delivery volumes must be reliable and adequate.
• After placement, dimensional stability must be achieved.
• Segregation should be minimized.

The most reliable measurement of backfill stability is strength, but the placed (in-situ)
quality also depends on:

• The effectiveness of the placement technique


• Segregation after placement (partly dependent on the placement technique)
• The consistency of the backfill mix (quality control)
• Post-filling water and solids loss (including any binder present)

3.9.5.5 Backfill placement


The placed backfill strength is dependent on the following:

• The delivered density


• The final density (after any water and solids loss and gravity settlement)
• The particle size distribution
• The binder type and content (if any)

Backfill is required in situ mainly under two basic placement schemes:

• Bulk placement where the fill is a working floor (platform) or roof. This placement
is relatively easy, and the only major potential problems are segregation of the fill
and binder loss in the drainage water.
• Tight placement where the fill is typically used as an artificial sidewall and needs
to have reliable contact with the roof of the stope. The potential problems are to
control/eliminate post-filling shrinkage and avoid costly containment structures
and/or mining at unfavourable gradients from a production standpoint.

References
Ankenman, B., Nelson, B.L. & Staum, J. (2010) Stochastic Kriging for simulation metamodeling.
Oper Res. Vol. 58, pp. 371–382.
Atlas Copco (1997) Guide to underground methods and applications.
Barton, N. (1976) Recent experiences with the Q-system. Proceedings of Symposium on
Exploration for Rock Engineering, Johannesburg, South Africa. January.
Barton, N., Lien, R. & Lunde, J. (1974) Engineering classification of rock masses for the design of
tunnel support. Rock Mech. Vol. 6, pp. 189–236.
Belem, T. & Benzaazoua, M. (2008) Design and application of underground mine paste backfill
technology. Geotech Geol Eng. Vol. 26, pp. 147–174.
340 Underground hard rock mining

Bieniawski, Z.T. (1973) Engineering classification of jointed rock masses. Trans S Afr Inst Civil
Eng. Vol. 15, pp. 335–344.
Bieniawski, Z.T. (1974) Estimating the strength of rock materials. J S Afr Inst Min & Metall.
Vol. 74, No. 8, pp. 312–320.
Bieniawski, Z.T. (1984) The design process in rock engineering. Rock Mechanics Design in Mining
and Tunnelling. AA Balkema Publishers.
Bieniawski, Z.T. (1989) Engineering Rock Mass Classifications: A Complete Manual for Engineers and
Geologists in Mining, Civil, and Petroleum Engineering. John Wiley & Sons, London, United Kingdom.
Brady, B.H.G & Brown, E.T. (1985) Rock Mechanics and Underground Mining. George Allen &
Unwin, Australia.
Brown, E. (2003) Underground Mining Methods.
Chamber of Mines Research Organization (COMRO) (1988) An Industry Guide to Methods of
Ameliorating the Hazards of Rockfalls and Rockbursts. Johannesburg, South Africa.
Cook, N.G.W. & Salamon, M.D.G. (1966) Rock mechanics applied to the study of rockbursts.
J S Afr Inst Min Metall. Vol. 66, No. 3, pp. 97–108.
Das Sharma, P. (2012) Weblog: Rock breakage and blast design considerations in openpit.
October 12. miningandblasting.wordpress.com/2012/10/12/rock-breakage-and-blast-design-
considerations-in-openpit/.
De La Vergne, J.N. (2003) The Hard Rock Miner’s Handbook, 3rd Edition, pp. 42–43. McIntosh
Engineering, USA.
De Souza, E. (2010) Room and Pillar. The Robert M. Buchan Department of Mining, Queens
University, Kingston, Ontario, Canada.
Deere, D.U. (1964) Technical description of rock cores for engineering purposes. Rock Mech Eng
Geol. Vol. 1, No. 1, pp. 16–22.
Deere, D.U. & Deere, D.W. (1988) The Rock Quality Designation (RQD) Index in Practice, ASTM
ST 984, Philadelphia, US.
Denkhaus, H.G. (1963) Critical review of strata movement theories and their application to prac-
tical problems. J S Afr Inst Min Metall. Vol. 63, No. 8, pp. 310–332.
Dominy, S.C., Noppé, M.A. & Annels, A.E. (2002) Errors and uncertainty in mineral resource
and ore reserve estimation: The importance of getting it right. Explor Min Geol. Vol. 11,
No. 1–4, pp. 77–98.
Environmental Systems Research Institute [ESRI] (2016) How Kriging Works. https://desktop.
arcgis.com/en/arcmap/10.3/ tools/3d-analyst-toolbox/how-kriging-works.htm.
Forbes, B., Vlachopoulos, N. & Hyett, A.J. (2018) The application of distributed optical strain sens-
ing to measure the strain distribution of ground support members. FACETS. Vol. 3, pp. 195–226.
Frimpong, S. (2011) Course Notes on Mine Planning and Design (Mi Eng 393). Missouri University
of Science and Technology, Rolla, MO, USA.
Gertsch, R.E & Bullock, R.L. (1998) Techniques in Underground Mining. SME, Colorado.
September.
Goldfields (2009) Mineral resources and mineral reserves. South Deep Gold Mine-Technical short
report. www.goldfields.com/reports/rr_2009/tech_south.php/tech_south.php.
Grimstad, E. & Barton, N. (1993) Updating of the Q-system for NMT. Proceedings of
International Symposium on Sprayed Concrete, Fagernes, Norway. pp. 46–66.
Hamrin, H. (1982) Choosing an underground mining method. In: Hustrulid, W.A. (ed)
Underground Mining Methods Handbook. AIME, New York, pp. 88–112.
Hamrin, H. (2001) Underground mining methods and applications. In: Hustrulid, W.A. &
Bullock, R.L. (eds) Underground Mining Methods: Engineering Fundamentals and International
Case Studies. Society for Mining Metallurgy and Exploration, Littleton, pp. 3–14.
Hassell, R., Villaescusa, E., de Vries, R. & Player, J. (2015) Stope Blast Vibration Analysis at
Dugald River Underground Mine. 11th International Symposium on Rock Fragmentation by
Blasting. August.
Hedrick, N. (2010) Yielding rock bolt. US patent 07645096. Garford Pty. Ltd. December 1.
Underground hard rock mining 341

Hoek, E. & Brown, E.T. (1980) Underground Excavations in Rock. Institution of Mining and
Metallurgy, London.
Hoek, E. & Wood, D.F. (1987) Support in underground hard rock mines. In: Udd, J. (ed) Underground
Support Systems. Canadian Institute of Mining and Metallurgy, Montreal. Special Volume 35.
Hoek, E., Kaiser, P.K. & Bawden, W.F. (1993) Support of Underground Excavations in Hard Rock.
Funded by Mining Research Directorate and Universities Research Initiative Fund, Canada.
Hormazabal, E., Alvarez, R., Russo, A. & Acevedo, D. (2018) Influence of the undercut height
on the behaviour of pillars at the extraction level in block and panel caving operations. Caving
2018, Australian Centre for Geomechanics.
Hustrulid, W.A. (1982) Underground Mining Methods Handbook. SME-AIME, New York.
Hustrulid, W.A. & Bullock, R.L. (2001) Underground Mining Methods: Engineering Fundamentals
and International Case Studies. Society for Mining Metallurgy and Exploration, Littleton.
Jager, A.J. (1992) Two new support units for the control of rockburst damage. Proc Int Symp on
Rock Support, Sudbury, Canada. pp. 621–631.
Jakubec, J. & Laubscher, D.H. (2000) The MRMR Rock Mass Classification System in Mining
Practice. MassMin, Brisbane, Australia.
Kempson, W.J., Webber-Youngman, R.C.W. & Meyer, J.P. (2015) Optimising shaft pressure losses
through computational fluid dynamic modelling. Appl Therm Eng. Vol. 90, pp. 1098–1108.
Kim, B.H., Cai, M., Kaiser, P.K. & Yang, H.S. (2007) Estimation of block sizes for rock masses
with nonpersistent joints. Rock Mech Rock Eng. Vol. 40, No. 2, pp. 169 –192.
Krige, D.G. (1951) A statistical approach to some basic mine valuation problems on the
Witwatersrand. J Chem Metall Min Soc S Afr. Vol. 52, pp. 119–139.
Lacy, W. (2015) An Introduction to Geology and Hard Rock Mining. Rocky Mountain Mineral Law
Foundation: Science & Technology Series.
Laubscher, D.H. (1975) Class distinction in rock masses. Coal Gold Base Miner S Afr. Vol. 23, August.
Laubscher, D.H. (1990) A geomechanics classification of jointed rock masses: Mining applica-
tions. Trans Instn Min Metall. pp. A1–A8.
Laubscher, D.H. (1994) Cave mining: The state of the art. J S Afr Inst Min Metall. pp. 279–293.
October.
Melbye T. (2006) Sprayed concrete for rock support. BASF Handbook, 11th Edition. BASF
Construction Chemicals, Zurich, Switzerland.
Michaud, D. (2016) Mining methods .911 Metallurgist. March. www.911metallurgist.com/blog/
mining-methods#shrinkage-stoping.
Mine Safety & Health Administration (MSHA) (2021) Safety alert: Electric vs electronic detona-
tors. January 29. www.msha.gov/ime-alliance-electronic-detonators-safety-alert.
Morelli, G.H. (2014) On joint roughness: Measurements and use in rock mass characterization.
Geotech Geol Eng. Vol. 32, pp. 345–362.
Normet (2006). D Bolt. www.normet.com/wp-content/uploads/2016/09/d-bolt-tds-global-20200511.pdf.
Ortlepp, W.D.O (1989) Guidelines for the Early Removal of the Shaft Reef Area. Anglo American
Corp of SA Ltd., Johannesburg (June).
Preston, C. (1995) 3D Blast design for ring blasting in underground mines. Proceedings of EXPLO
95 – The Australasian Institute of Mining and Metallurgy, Brisbane, Australia.
Pugh, G.M. & Rasmussen, D.G. (1998) Cost calculations for underhand cut-and-fill mining.
Techniques in Underground Mining, Ch 34. SME, Colorado. pp. 595–601.
Queens University (2011a) Cut and fill. April. www.minewiki.engineering.queensu.ca/mediawiki/
index.php/Cut_and_fill.
Queens University (2011b) Sub-level open-stoping. April. www.minewiki.engineering.queensu.
ca/mediawiki/index.php/Sub-level_open_stoping.
Queens University (2012) Mine design: Sub-level caving. January. www.minewiki.engineering.
queensu.ca/mediawiki/index.php/Sublevel_caving.
Queens University (2014) Vertical Crater Retreat. February. www.minewiki.engineering.queensu.
ca/mediawiki/index.php/Vertical_crater_retreat.
342 Underground hard rock mining

Raffaldi, M.J., Seymour, J.B., Abraham, H., Zahl, E. & Board, M. (2018) Cemented paste backfill
geomechanics at the Lucky Friday Mine. ARMA 18-815 (June).
Rodriguez, R. (2017) Anatomy of a Mine. National University of Columbia, Columbia. www.
medellin.unal.edu.co/~rrodriguez/geologia/anatomy-of-a-mine/Anatomy.
Rowland, J.H. & Mainiero, R. (2000) Factors affecting ANFO fumes production. NIOSH TIC2
Number: 20021058. February.
Salamon, M.D.G. (1974) Rock mechanics of underground excavations. Advances in rock mechan-
ics. Proceedings of 3rd Congress of the International Society of Rock Mechanics, Denver,
USA. Vol. 1b, pp. 91–109.
Sauer, G. & Gold, H. (1989) NATM Ground support concepts and their effect on contract prac-
tices. Rapid Excavation and Tunnelling Conference, Los Angeles, US. June.
Shi, G.H. (1988) Discontinuous deformation analysis: A new numerical model for the statics and
dynamics of block systems. PhD Thesis. University of California, Berkeley.
Spearing, A.J.S. (1989) The modification of Witwatersrand gold plant tailings for optimum appli-
cation as an underground support medium. Thesis for MS in Civil Engineering. University of
the Witwatersrand, Johannesburg, South Africa. December.
Spearing, A.J.S. (1995) Handbook on Hard-Rock Strata Control. SAIMM, ISBN 1 874832-45-5.
Spearing, A.J.S. (2002) Shotcrete (sprayed concrete) for mining applications. Released by the
Underground Construction Group, Degussa Chemicals.
Stillborg, B. (1986) Professional User’s Handbook for Rockbolting. Trans Tech Publications,
Clausthal-Zellerfeld.
Tucker, H., Holder, A., Swarts, B., van Strijp, T. & Grobler, E. (2016) The CCUT block cave
design for Cullinan Diamond Mine. J S Afr Inst Min Metall. Vol. 116, No. 8.
University of Wollongong (2020) Theories of rock bolting. www.miningst.com/rock-bolting/theories-
of-rock-bolting/.
Vandewalle, M. (1998) The use of steel fibre reinforced shotcrete for the support of mine open-
ings. J.S.Afr. Inst Min Metall Vol. 98. May/June. pp. 113–120.
Vardan, H. & Murthy, S.N. (2007) An experimental investigation of jack hammer drill noise with
special emphasis on drilling in rocks of different compressive strengths. Noise Control Eng J.
Vol. 55, No. 3. pp 282–293.
Warner, A.W. (1934) Progress in low ranks of coal. Ind Eng Chem. Vol. 26, No. 2, pp. 54–164.
Wagner, H. (1980) Pillar design in coal mines. J S Afr Inst Min Metall. Vol. 80, No. 1, pp. 37–45.
Williams, T.J., Brady, T.M., Bayer, D.C, Bren, M.J., Paklanis, R.T., Marjerison, J.A. & Langston,
R.B. (2007) Underhand cut and fill mining as practiced in three deep hard rock mines in the
United States. NIOSH, CDC NIOSHTIC2 Number: 20033309. April.
Wilson, J.W. & More O’Ferrall, R.C. (1970) The application of the electrical analogue to mining
operations. J S Afr Inst Min Metall. Vol. 70, No. 6, pp. 115–148.
Wiseman, N. (1979) Factors effecting the design and conditions of mine tunnels. Research Report
No. G01G10. Chamber of Mines Research Organization, Johannesburg, South Africa.
Worsey, P.N. (2001) Blasting Design and Technology Lecture Series. University of Missouri-Rolla,
Rolla, USA.
Zhang, Q. & Wang, X. (2016) Underground Mining Technics of Metallic Ore Deposits. University
of Central South Press, Changsha, China.
Zhang, Z.X. (2008) Impact of rock blasting on rock engineering. MassMin 2008, Proceedings of 5th
International Conference & Exhibition on Mass Mining, Luleå, Sweden, June 9–11. pp. 671–680.
Zhou, J., Shi, X. & Li, X. (2015) Utilizing gradient boosted machine for the prediction of damage
to residential structures owing to blasting vibrations of open pit mining. Journal of Vibration &
Control. February pp. 1–12. DOI: 10.1177/1077546314568172.
Chapter 4

Green and sustainable mining

4.1 Green mining

4.1.1 The continued need for mining


Civilization has been defined by mining and its impact on improving human stand-
ards of living. Our development eras have been named after the mining material that
has had the most influence on mankind’s development in the era (Age):

• Stone Age – pre 4000 bc


• Bronze Age – 4000 bc to 1500 bc
• Iron Age – 1500 bc to about ad 1800
• Steel Age – ad 1800 to about 1945
• Nuclear Age – After 1945

An example of the importance of mining is shown in Figure 4.1, which is based on


specific data for the US, but it would be similar in most countries.
Mining will continue to be needed in the future as it has been essential in the past,
but the minerals in critical demand will change as they have been changing in the past.
Consider how the need for thermal coal, lead and asbestos has been reducing but the
need for rare earths and other modern battery materials is increasing.
Mining in the future will continue to be essential like the need for energy (in which
mining is also an essential direct part), food and clean water. There are no replace-
ments for the building materials, fertilizers and farm implements, communication sys-
tems, transport and even mineral health supplements to name a few.
The way mining is undertaken will and must continue to evolve and improve.
Aspects such as environmental protection, mine safety and the social license to oper-
ate will become focal points especially for new mining ventures.

4.1.2 United Nations Sustainable Development Goals


The United Nations (UN) developed Sustainable Development Goals (SDGs) and
published them in 2015 to be “a blueprint to achieve a better and more sustainable
future for all”. This was a global and necessary first step to try and respond to seri-
ous environmental, economic imbalances and political problems that affected all of
mankind.

DOI: 10.1201/9781003185680-4
344 Green and sustainable mining

Figure 4.1 The need for mining.

According to the UN, these replaced the earlier Millennium Development Goals
(MDGs) that were introduced in 2000 with the main goal of fighting poverty.
According to the UN, the main quantifiable benefits from implementing the MDGs
were (it should be noted that some actions started earlier than the MDGs’ adoption):

• More than 1 billion people have been lifted out of extreme poverty (since 1990).
• Child mortality dropped by more than half (since 1990).
• The number of out-of-school children has dropped by more than half (since 1990).
• HIV/AIDS infections fell by almost 40% (since 2000).

The UN identified 17 key areas under its SDGs as shown in Figure 4.2.
Mining directly has an influence over many of them but especially the following:

• #6 – Clean Water and Sanitation


• #7 – Affordable and Clean Energy
• #8 – Decent Work and Economic Growth
• #9 – Industry, Innovation and Infrastructure
• #12 – Responsible Consumption and Production
• #13 – Climate Action
• #15 – Life on Land

Mining also indirectly affects most of the other goals too. The mining industry there-
fore will need to be a key player going forward if the global SDGs are to be achieved.
This is especially relevant as strategic mining of certain metals and minerals, such as
battery minerals, will need to increase dramatically in order to reach greenhouse gas
reduction future targets.

4.1.3 The need for green mining


For all the positives associated with the mining industry, mining has also had an
adverse impact on the environment mainly by damaging the environment with huge
amounts of waste disposed of on surface and polluting (and damaging) surface
Green and sustainable mining 345

Figure 4.2 UN Sustainable development goals

and underground water. Safety for mine employees continues to improve, but with
few exceptions, the mining industry globally has still a long way to go to achieve
near “zero harm”, although there are many positive developments in this. Consider
the remarkable safety achievement of the platinum mine in South Africa shown in
Figure 4.3.
Probably the biggest environmental impact caused by mining is from old, previ-
ously mined areas that are contaminated and uncontrolled. It is only recently that
most countries have adopted adequate legislation to protect the environment after
mining ceases. People live in these historically contaminated areas and their health
can be adversely affected.

Figure 4.3 Fatality achievement in South Africa


Source: Media release from Impala Rustenberg Services, September 2021
346 Green and sustainable mining

The other major concern is waste tailings dams. The problems with conventional
tailings dams include the following:

• The inherent safety risk of liquefaction such as what recently occurred in Brazil
(2019), causing significant loss of life, extreme environmental damage and severe
property damage. This is clearly the most significant and poses the largest hazard
and represents the area that needs to be addressed immediately.
• Serious water losses from evaporation on the dams in regions where water is a
scarce commodity (such as in Australia and South Africa).
• The large area needed for the dam and the cost of eventual reclamation and often
on-going treatment.

A recent major tailings dam disaster occurred at Corrego do Feijao iron ore mine owned
by Vale, in Brumadinho, Brazil, in January 2019 (Wikipedia, 2019). This was the mining
companies’ second multiple fatality tailings dam collapse in 3 years. In 2015 at Samarco,
a jointly owned mine with BHP, a tailings dam collapsed, killing 19 people, in Minas
Gerais, Brazil. The result was 270 people dead and about 12 million m³ of tailings
released into the environment. A company-wide dam audit by the company in April
2019 resulted in a further 18 tailings dams or retaining dykes that could not be con-
firmed as stable. Such failures have got to become a thing of the past, and selective
processing and backfill are a couple of ways through which the waste on surface can be
reduced significantly.
According to MIT’s “Mission 2016”, there is potential to choose more environmen-
tally friendly mining processes. Another broad method for improving efficiency would
be to address the general mining process and purification processes.
Although openpit mining contributes about 85% of all mineral mining, it is one
of the most environmentally taxing. About 73% of extracted rock goes to waste.
Meanwhile, underground mining wastes is only 7% of the global extracted rock and
is more expensive to produce (Hartmann & Mutmansky, 2002). In-situ mining can be
more environmentally friendly than underground mining and is cheaper than many
mining methods (Ulmer-Scholle, 2022). However, in-situ mining cannot be imple-
mented in all cases as the ore needs to be beneath the water table (the level at which the
ground is saturated with water) and it needs to be porous enough to let the leaching
solution dissolve (Topf, 2011). Unfortunately, in-situ leaching can also be very harmful
if the solution leaks into the water supply. There are plenty of examples of past leaks at
in-situ leaching mines (“Coloradoans Against Resource Destruction”, 2008).
It is unfeasible to convert all current mines to more environmentally friendly min-
ing methods due to economic constraints and ore deposit geography. However, when
opening new mines in areas with low risks of water contamination, in-situ leaching
should be the choice method when physically possible. If not, then the environmen-
tal benefits of underground mining need to be weighed with the financial benefits of
openpit mining to determine the mining method of choice on a mine-by-mine basis.
Mining, moving forward, will therefore need to effectively address, undertake and
implement the following (World Economic Forum, 2020):

• Transition to a low carbon economy – To control climate change, it is essential that


mining focusses on more energy efficiency and providing materials vital for more
environmentally friendly energy production.
Green and sustainable mining 347

• Access to resources – Easier and higher grade orebodies are being depleted so
effort is needed to locate and efficiently mine more difficult orebodies possibly at
a lower grade. This will require embracing new technologies.
• New ways to finance mining – Financing for new and higher risk orebodies
needs novel ways to spread the financial risk away from only the mining com-
pany. Successful methods have included royalty and metal stream supply
agreements.
• A social contract for mining – Local communities must obtain real and long-term
benefits from mining in their locations.
• Big data for mining – Real-time monitoring, robotics, artificial intelligence and
automation have the proven potential to not only improve safety but also produc-
tivity in the mining industry.
• The geopolitics of mining – Geopolitical risk and protectionism are on the rise
particularly as countries try to protect their reserves of critical minerals for their
own consumption.
• Modern mining workforces – Training and upskilling the workforce will be
essential and will need to be done in an open and collaborative manner with all
involved.

Figure 4.4 shows the future goals and objectives for mining according to the World
Economic Forum (2019).

4.1.4 The concepts of green mining


This involves combining sustainability and the environment as these are the basis for
all green mining.
Kirkey (2014) states: “Green mining is defined as technologies, best practices and
mine processes that are implemented as a means to reduce the environmental impacts
associated with the extraction and processing of metals and minerals”.
According to MIT (Mission 2016):
The plan for improving efficiency and decreasing the environmental impact of min-
ing is broken up into the following categories:

• Shutting down illegal and unregulated mines


• Choosing environmentally friendly general mining processes
• Implementing recently discovered green mining technologies
• Cleaning up the sites of shut-down mines
• Re-evaluating cut-off grades
• Research and development of green mining technology

The bottom line for mining is:

Irrespective of the range of possible scenarios, conclusions can be drawn for


the industry: Mining will not disappear; primary extraction will continue but
volumes are unlikely to grow in line with GDP growth. This means that pres-
sure to realize scale effects and cost efficiency will remain in the foreseeable
future. Demands for cost-effectiveness will exist in parallel with demand for
348 Green and sustainable mining

Figure 4.4 The future goals and objectives for mining (World Economic Forum)

environmentally and socially responsible actions, leading to new partnership


and operating models.
World Economic Forum

Modern mining is becoming more environmentally responsible and the goals of green
mining are basically:

• Small footprint (low emissions)


• No or minimal environmental disruption
• Safe mining
• Sustainable mining (social license to operate)

Mining sustainability in this context means that:

• During mine exploration, development, exploitation and final rehabilitation,


local communities (especially First Nation) are consulted, respected and their
traditions protected.
Green and sustainable mining 349

• After mining, the surface is fully reclaimed.


• Local communities are left economically viable.

According to Prof. John Meech, Director of the Centre for Environmental Research
into Minerals, Metals and Materials in Canada, sustainable mining is a combination of:

• Techno-economic issues
• Environmental issues
• Socio-political issues

The following is taken from Colwell (2017):

It is estimated that between now and 2050, the world’s population will grow from
just over 7 billion to 9.6 billion, along with further growth in consumption per
capita. The bulk of that growth will take place in Africa and Asia, where the
demand for an improved quality of life will drive the need to access goods and ser-
vices. Africa, for example, now has the fastest-growing middle class in the world.
Some 313 million people, 34% of Africa’s population, spend $2–20 a day, a 100%
rise in less than 20 years, according to the African Development Bank.
Success in sustainability means looking beyond the perspectives of material
exhaustion, rising costs or social injustice. The integrated view also pays special
attention to the role that strategy, technology, policies, preferences and vari-
ous stakeholders will play in the future availability of primary and secondary
resources. Today, we’re focusing on where and how the circular economy is forc-
ing change in the metals and mining industry. Why, you ask? Because … Mining
and metals are essential to global economic and social development, and are con-
nected to almost all industry value chains.

Waste is a critical issue along the whole metals value chain, from mining waste to
eventual end-of-life products such as scrap steel from construction and demolition
waste or the growing problem of electronic waste.
Being one of the world’s largest waste generators, the mining sector can adopt sim-
ilar logic to that of the circular economy to improve its sustainability performance.
The baseline shows that mining and metals are among the world’s great generators of
waste, accounting for around 10 billion tonnes a year, around 40–55% of the global total.
Tailings, the waste from extractive processes excluding overburden, can hold large
potential value: Current estimates suggest that with the right technology for treating
bauxite waste, aluminium production could be increased by 20%, a potentially huge
capture of value from tailings.
The mining value chain is characterized as a linear process that generates large
volumes of waste.
Each waste stream along the metal value chain has its own set of environmental issues.
In many cases, the most challenging (and massive) waste stream within the metals
supply chain is upstream (i.e., mining waste).
Due to significantly lower grades for most extracted minerals and metals, tailings
can account for up to 99% of crushed and ground ores. In addition, there is also a
“hidden” unaccounted flow of waste rock and overburden. Different strategies for
managing mining waste can be characterized in terms of their ability to decrease the
350 Green and sustainable mining

risks and consequences of environmental legacy and in generating economic value out
of waste.
An integrated, multidisciplinary approach to mining and metal waste is needed:
Without such an approach, it will not be possible to account for the different social,
economic and environmental dimensions of sustainability, engage with the network
of actors within the metal supply chain and look beyond short-term economic ben-
efits and risk-averse behaviour to target the supply of restorative and regenerative
resources in a circular economy model.
Data from World Mining Data (2020) shows the global mining annual production
to have increased as follows:

• 1985: 9.6 billion tonnes


• 2000: 11.3 billion tonnes
• 2018: 17.7 billion tonnes

The waste produced from this is huge and will not be sustainable in the future.
In its current form, it looks like we may need more than two planets’ worth of mate-
rials by 2050.
In a world with increasingly constrained resources and many environmental chal-
lenges, the balance of supply and demand will shift for key commodities. This will
have fundamental impacts across the mining and metals value chain.
Effective green mining initiatives that have potential include:

• The active involvement of the local community


• The potential for in-situ remote mining and leaching
• More real-time monitoring of all active and passive mining activities
• More selective mining (grade control)
• The utilization of renewable energy in mines
• More water recycling and better treatment before excess water is released from
the mine
• More effective mine waste treatment and disposal
• More efficient methane capture and utilization on coal mines
• Replacing diesel-powered vehicles
• The introduction of more environmentally friendly extraction systems
• Better training using virtual reality to make it more effective
• More backfilling on mines and more effective surface tailings facilities with max-
imum repurposing

4.1.5 The Fourth Industrial Revolution and its impact on mining


The industrial revolutions that have shaped the major development of our world have been:

• The First Industrial Revolution occurred between the late 1700s and 1800s and
marked the move from hand production to the first mechanized production with
the advent of the steam engine.
• The Second Industrial Revolution occurred from the late 1800s to the early 1900s
and saw the development of electrical grids, improved communication (using the
telegraph) and more rapid and reliable transportation (mainly rail).
Green and sustainable mining 351

Figure 4.5 The relationship between data and understanding

• The Third Industrial Revolution (the Digital Revolution) started in the mid-1900s
and saw the development of computers.

The Fourth Industrial Revolution is the automation of traditional manufacturing and


industrial processes. Schwab (2017) predicted the revolution to focus on developments
in big data-based technologies such as: robotics, artificial intelligence, nanotechnology,
the Internet of Things (IoT), 3D printing and fully autonomous vehicles amongst
other developments.
The Fifth Industrial Revolution, which has already started, can be simplified as
humans and robots (machines) working together in harmony.
The mining industry has the most to gain from the Fourth Industrial Revolution as
mining can benefit the most from real-time monitoring and automation since it is not
conducted in engineered materials and deals with nature, which is essentially random
and unpredictable.
Most mines especially in the exploration and development stage lack data and
understanding and reside in Region 3 in Figure 4.5. In addition, most mining and
process operations currently experience a serious lag between identifying a problem
and being able to take an action to correct it, which leads to inefficiencies. Real-time
monitoring, artificial intelligence, automation and robotics have the potential to dra-
matically improve mine safety and efficiencies if carefully and thoroughly tested and
introduced.

4.2 The circular mining economy

4.2.1 The current mining industry and the need for change
Mining is key to the global economy and our standard of living.
Mining is becoming more challenging, however, as higher grade and easier mined
deposits are being depleted and new deposits tend to be in more harsh climates, are
deeper and with lower grades, which makes it more risky.
According to Price Waterhouse Cooper (PWC Mining, 2019), although the top 40
global mining companies have performed very well financially, investors are con-
cerned whether mining can create sustainable and environmentally friendly growth
in the future. Climate change concerns and how the mining industry can become part
of the solution has created some doubt amongst investors. High-profile environmental
disasters such as the iron ore waste dam collapse in Brazil in 2019 have also had a very
negative effect on the public’s perception of the mining industry as a whole.
352 Green and sustainable mining

According to the PWC Mining 2019 report:

The mining industry will have a window of opportunity over the next few years,
created by strong operating fundamentals, to adapt to the growing and changing
expectations of stakeholders. By utilising technology to operate safely and more
efficiently, addressing global concerns, and maintaining a disciplined strategy to
create ongoing value for its stakeholders, the industry can forge a better future
for all beneficiaries of mining – industry, consumers, communities and other
stakeholders.

Previously, mines have operated under the specific country’s umbrella that granted
the mining licenses required and their concerns were the workforce and shareholders’
return on investment (ROI). These two legs underneath the umbrella have now grown
to include a third and equally important leg: looking after the needs and aspirations
of the local community. This includes:

• Water protection
• The environment
• A social license to operate, which includes the local community benefitting from
the mining operation directly (during and after exploitation)

Globally the mining industry needs to also address the vast waste that it generates and
the power that it consumes especially in comminuting (crushing and grinding) ore.
Compared with other industries, such as energy, mining has been slow overall
to implement new digital technologies, with all the advantages if implemented
correctly.

4.2.2 The need for a circular mining economy


The key tenets of the circular economy are utilizing resources efficiently, limiting final
waste disposal and reducing losses of material with direct or indirect value. It is basi-
cally a concept to design waste out of the economy, as is needed as resources tend to
shrink whilst the global population continues to increase.
A circular economy is basically an environmentally friendly economy as shown in
Figure 4.6. The circular economy is a closed loop of resource-product-recycle, with the
minimum of waste. The goal is to achieve a balance between minimal new production
and consumption with as little waste, and maximum reuse, repurpose and recycle.
It is generally agreed that the principle a circular economy should follow is “Reduce,
Reuse, Recycle”, or “3Rs” for short. These mean the following:

• “Reduce” refers to reducing the flow of new material and energy into the produc-
tion and consumption process.
• “Reuse” is associated with the procedure method with the aim of extending the
useful life of a product or service (repairing rather than disposing).
• “Recycle” belongs to the output method, meaning that materials are returned to
renewable resources after making use of it.

The concept of the circular economy has been gaining traction particularly in both
Europe and China (McDowall et al., 2017). The circular economy decouples growth
Green and sustainable mining 353

Figure 4.6 The outline of a circular mining economy


Source: EIT Raw Materials, www.eitrawmaterials.eu

from natural resource consumption. The circular economy (along with other related
sustainability concepts) provides a system perspective of waste elimination through
the rethinking and redesign of products and processes along the value chain and
between supplier networks.
Lacy and Rutqvist (2015) demonstrated that a circular economy could deliver
$4.5 trillion in revenue to a range of companies by as early as 2030.
One approach to accelerate the uptake of the circular economy is by introducing
new innovative business models that deeply embed its principles into the way that
companies generate and capture economic value.
Circular business models are disruptive and require innovative business models that
aim to drive the sustainability of the whole business network through circularity.
In the words of Unilever CEO Paul Polman, “[a circular] economy can deliver
growth. Innovative product designers and business leaders are already venturing into
this space”.
With a potential $1 trillion opportunity in transitioning to the circular economy,
companies are recognizing that preservation makes as much economic sense as it does
to the environment.
Although it still seems that the majority of global debate is focussed on energy,
resource sustainability is also rapidly becoming an area of focus.
The pace of regulation is increasing, for example, with the introduction of the
Circular Economy Package in the European Union (EU) starting in 2012, posing
many threats and opportunities for miners (European Commission, 2012).
354 Green and sustainable mining

The European Resource Efficiency Platform (EREP) adopted by the EU in December


2012 states:
The EU will need a systemic change in the use and recovery of resources in order
to improve the resilience of our environment, societies and economies, within the
boundaries of the planet. This should boost competitiveness and contribute to a
sustainable, reindustrialised European economy. According to a recent estimate,
the EU could realistically reduce the total material requirements of its economy
by 17% to 24%, boosting GDP and creating between 1.4 and 2.8 million jobs.
To respond to the key policy challenges at hand, our recommendations are
designed to:
• Create growth and jobs.
• Provide incentives to overcome barriers to improving resource efficiency.
• Put a proper value on resources.
• Provide clear information and measure progress.
• Promote new business models.
Targets are essential for guiding action, for making sure that we are moving in the
right direction, while indicators are needed to measure progress.

The circular economy is challenging the mining industry to improve sustainability by


repurposing and generating some value from mining waste, possibly as a backfill for
mines or as a raw material possibly in the construction and manufacturing industries.
Mining faces a difficult challenge to achieve sustainability and the social license to
operate, but it must be achieved and most mines are working towards it. Better coor-
dination and communication across the mining industry would help.
The nature of mining is connected with the depletion of mineral-rich resources, as
massive amounts of rock are processed to extract the metals and minerals that are
needed in ever-increasing amounts by the modern world.
Sustainability in the mining industry is complex and involves all the activities from
exploration, exploitation, social and environmental issues and reclamation.
Metals and mining companies need to adopt a fresh perspective on what it means to
create value, or better a shared value.
While the mining industry has only made a limited contribution to the circular
economy so far, current market conditions, which are prompting calls for greater
innovation, make the timing right for the industry to boost its contribution by utiliz-
ing and generating value from mining waste.
The opportunity to create value and reduce environmental liability from waste
streams along the value chain is potentially one way that the mining and metals indus-
try could make substantial contributions to the circular economy and, in doing so,
improve sustainable development.
According to Lacy and Rutqvist (2015), there are five business models that compa-
nies should consider across their value chain:
• Circular supply chain
• Recovery and recycling
• Product life extension
• Sharing platforms
• Product as a service
Green and sustainable mining 355

These will influence how miners source materials and energy, optimize their opera-
tional footprint and make them consider where to play in the value chain and how to
partner with customers.
For miners, circular economy winners will be equally focused on managing
resources in the market and digging to extract additional resources where needed. The
trick is to understand how the supply/demand trends play out for each commodity, to
identify the relevant risks and opportunities and to position for future growth before
the circular economy becomes a reality. There are various opportunities to implement
circular flows at the mine site level, which would result in enhancing mineral extrac-
tion, reducing mineral losses to mining waste and mitigating some of the environmen-
tal impacts related to mine waste disposal.
Corder (2021) has undertaken research in Australia and has produced a framework
that allows a mine site’s performance to be assessed with regards to the circular flow of
minerals. He states that modelling business practices around a circular economy will
see the industry become more sustainable, reduce costs to the environment and enable
companies to manage the cradle-to-grave operations.
Raising the comprehensive benefit of resource development by reducing emissions
of various pollutants such as tailings, gangue and mine wastewater.

4.2.3 Achieving a circular economy


To achieve a circular economy, the key requirements are therefore (modified from
PWC, 2019).
Reuse: A product’s, or a component’s, life cycle should be prolonged by reuse.
For this to be possible, products need to be redesigned and standardized by
manufacturers so that components can be easily and cost effectively extracted for
reuse.
Modular and standardized construction: For this to work, standardization of com-
ponent design and specifications and close collaboration and cooperation along the
value chain are needed.
Reuse can also be achieved through re-manufacturing and refurbishing: this
includes the disassembly, cleaning, repair and reassembly of a product – restoring it
to a like-new condition.
Recycle: This requires reducing waste by reprocessing the mineral resources
products that have completed their initial functions so that they become available
resources again and can enter the secondary market.
Downstream recycling (old scrap): reclaim metal from products which have reached
their end of life or end of use. There are four prerequisites for downstream recycling
to be sustainable and efficient:

• Adequate collection and pre-processing infrastructures


• Enough old scrap available for the downstream process facility to operate effi-
ciently (volumes depend on the lifespan of metal currently in use)
• Competitive production costs, because recycling competes with primary metal
production
• Possibility of recycling (including up-cycling) or reuse in different applications
(recycled materials cannot always be reused in high-quality applications due to
alloying or impurities)
356 Green and sustainable mining

Recycling requires three types of infrastructure:

• The collection and transportation of recyclable material


• Separating and sorting facilities to isolate components
• Reprocessing facilities, such as smelters and refining, to make new metals from
scrap

The geographic mismatch between places where recyclable material is available and
those where it is most needed represents a serious gap that must be resolved to achieve
a sustainable world. Most of the recyclable metal waste is created in industrialized and
high-income countries, while the countries which will need sufficient scrap to make
the transition from mining to a circular use of resources are mostly low-income and
have immature recycling sectors.

4.2.4 The challenges to creating a circular mining economy


Whilst a circular mining economy has the obvious advantages of ensuring adequate
raw materials for a growing global population, reducing environmental damage and
waste, it faces significant challenges.
According to Bartels (2019), the challenges are for mining and metals companies to:

• Safeguard their market share whilst also tapping into new value streams
• Adapt fast, embrace change and stay relevant
• Learn about proactive steps to capitalize on emerging opportunities and tap into
sustainable new revenue streams in the circular economy

She also notes that:

• 40% of steel produced is now made from scrap metal.


• 41 mobile phones can yield as much gold as one tonne of ore.
• $4.5 trillion of opportunity from eliminating waste is through the circular
economy.

Bartels further recommends that mining companies start by:

• Developing circular operations – Start by accelerating circular initiatives across


your mining and metals operations, for example, using real-time monitoring, ana-
lytics and predictive maintenance while promoting remanufacturing and end-of-
life recycling.
• Innovate new circular products and services – Secondly, engage with downstream
users of your materials to co-create innovative circular products and services,
enabled by advanced technologies to facilitate better recovery, reprocessing and
reuse.
• Collaborate with customers and build a circular partners’ ecosystem – Finally, col-
laborate proactively up and down your supply chains to drive industry momen-
tum and create a more favourable environment to retain ownership and benefit
from extended product lifecycles and improved circularity.
Green and sustainable mining 357

Other non-traditional mining companies are, however, entering the circular mining
economy and traditional mining companies need to formalize how they plan to com-
pete in this new emerging economy or risk losing relevance and profitability.
It is difficult to see how a circular mining economy can be realized without some form
of global association with powers to regulate new minerals supply to match what is needed
from new mined sources as recycling increases. Many could see this as a form of monop-
oly but unless all stakeholders are involved and coordinated, it is difficult to see how nor-
mal “supply and demand” would work. A large metal mine, for example, would want to
maximize output for economy of scale and would possibly undercut prices to have the
recycled metal less competitive, thus defeating the circular mining economy ideal.
Many may find this suggestion terrible as it would create a quasi-monopoly, but
perhaps a broader based organization, similar to the Organization of Petroleum
Exporting Countries (OPEC), OPEC + members and observers, may be able to create
a circular mining economy relatively quickly and the environmental and economic
global benefits could easily outweigh the downsides.

4.3 Sustainable development of minerals

4.3.1 Concepts and key factors


The United Nations Educational, Scientific and Cultural Organization (UNESCO)
defines sustainability as “development that meets the needs of the present without
compromising the ability of future generations to meet their own needs.”
According to Kuchling (2019), there are basically two views of what sustainability
in mining means:

• The first is general, and focusses on the ability of the mining industry to meet
future commodity needs and demands. It is concerned with keeping the mining
industry viable.
• The second is more related to stakeholders, with an emphasis on the benefits
a mining project will have on the local community in the short and long term
beyond the mining project’s life.

Whilst these are achievable goals, the general interested public and especially the local
communities believe that it needs to include the whole mining cycle, from processing
and subsequent use and including recycling, reuse and repurposing.
At a minimum, the key factors would seem to include:

• Ensuring the safety of the workforce and local community, during and after mining
• Minimizing the mine footprint
• Limiting damage to water and land resources
• Effective and environmentally sound mine closure
• Providing a benefit (usually economic) to the local community during and after mining

To achieve these, positive engagement by the mine with the local community needs to
happen early in the mining project, ideally at the exploration stage and continue trans-
parently until the successful closure of the mine. It means that the way that each new
mining project moves forward will be unique based on the specific local community’s
needs and aspirations.
358 Green and sustainable mining

4.3.2 Environmental impact assessments


Environmental impact assessments (EIAs) have become a major requirement for the
mine permitting process in many countries. If this is properly and thoroughly under-
taken, it can be very beneficial to the local community and have little if any adverse
effect on the environment.

4.3.2.1 Introduction
An environmental assessment (EA) is an assessment of the environmental con-
sequences (both positive and negative) of a proposed mining operation. It is
required in most countries before a mining permit is awarded and is therefore very
important.
The EIA is usually used when applied to projects by individuals or companies and
lists the actual environmental impacts that the proposed project could cause and out-
lines what actions will be taken to mitigate or eliminate these impacts. The term stra-
tegic environmental assessment (SEA) is sometimes used, and it applies to policies and
programs usually required by a specific country.
An EA is therefore a tool outlining the proposed environmental management to be
employed during the mining development, operation and closure. EAs may be gov-
erned by rules of administrative procedure regarding public participation and docu-
mentation of decision making, and may be subject to official review.
The purpose of the assessment is to ensure that the owners of a potential new mine
consider all the environmental impacts and their report can be reviewed by the rele-
vant authority as well as the public in a particular country.
EAs are rigorous, technical and have their own set of criteria. Independent techni-
cal experts should be brought in to ensure completeness and reduce partiality. There
are competent environmental consulting firms familiar with all aspects of a mining
EIA in most countries.
The costs of EAs vary depending on the size of a project but are normally less than
1% of the total cost of the project and are covered by the mining project.
The City of London, Canada, website states that an EIS is used to provide a suffi-
cient level of detail to demonstrate that a proposed development will have no negative
impacts on the natural features or ecological functions of the subject and surround-
ing (“adjacent”) lands. An EIS does not ensure that development proposals will be
approved. Their purpose is to inform the design and configuration of the develop-
ment, to avoid negative impacts at the outset, and to identify appropriate mitigation
and/or compensation for unavoidable impacts.
The International Association for Impact Assessment (IAIA) defines an EIA as “the
process of identifying, predicting, evaluating and mitigating the biophysical, social,
and other relevant effects of development proposals prior to major decisions being
taken and commitments made”. EIAs are unique in that they do not require adherence
to a predetermined environmental outcome, but rather they require decision makers
to account for environmental values in their decisions and to justify those decisions
in light of detailed environmental studies and public comments on the potential envi-
ronmental impacts.
Green and sustainable mining 359

4.3.2.2 The impact of mining


Mining requires large areas of land to be temporarily disturbed for the life of the
mine. This creates environmental challenges that can include:

• Surface subsidence and landslides


• Soil erosion
• Dust
• Noise
• Water pollution
• Biodiversity damage

Steps are taken in mining operations to minimize impacts on all aspects related
to the environment. This needs responsible pre-planning of all related operations,
implementing effective pollution control measures, monitoring and repairing dam-
age caused by mining and rehabilitating mined areas to at least the pre-mining
condition.
Aspects that are commonly investigated and addressed in an EIS are:
LAND DISTURBANCE

In best practice, studies of the immediate environment are carried out several years
before a coal mine opens in order to define the existing conditions and to identify
potential problems. The studies look at the impact of mining on surface and ground
water, soils, local land use, native vegetation and wildlife populations. Computer sim-
ulations can be undertaken to model impacts on the local environment. The findings
are then reviewed as part of the process leading to the award of a mining permit by the
relevant government authorities.
MINE SUBSIDENCE

Mine subsidence can be a problem with underground coal mining especially when
longwalling, because the ground surface lowers as a result of coal having been mined
beneath. A thorough understanding of subsidence patterns in a particular region allows
the effects of underground mining on the surface to be quantified. The coal mining
industry uses a range of engineering techniques to design the layout and dimensions of
its underground mine workings so that surface subsidence can be anticipated and con-
trolled. This ensures the safe, maximum recovery of a coal resource, while providing
protection to other land uses. Research in Solid Coal Backfilling undertaken by the
School of Mines at CUMT has virtually eliminated surface subsidence and its adverse
effects on surface and underground water.
WATER POLLUTION

Mine operations continuously work to improve water management, aiming to


reduce demand through efficiency, technology and the use of lower quality and recy-
cled water. Water pollution is generally controlled by carefully separating the water
runoff from undisturbed areas from water which contains sediments or salt from
mine workings. Clean runoff can be discharged into surrounding water courses,
while other water is treated and can be reused such as for dust suppression and in
coal preparation plants.
360 Green and sustainable mining

BIODIVERSITY

According to Sonter et al. (2018), mining poses serious and highly specific threats to
biodiversity. However, mining can also be a means for financing alternative livelihood
paths that, over the long term, may prevent biodiversity loss. In many regions, the con-
servation community cannot achieve biodiversity goals without engaging the mining
industry, yet few examples of effective collaboration currently exist. Mining companies
have financial incentive to mitigate biodiversity losses caused by their operations. This
needs to be addressed in an EIS.

• Dust reduction
• Noise control
• Social impact of mining

The Guidebook for Evaluating Mining Project EIAs, published by Environmental Law
Alliance Worldwide (2010), states that the benefits of an EIS are to:

• Potentially screen out environmentally unsound projects


• Propose modified designs to reduce environmental impacts
• Identify feasible alternatives
• Predict significant adverse impacts
• Identify mitigation measures to reduce, offset or eliminate major impacts
• Engage and inform potentially affected communities and individuals
• Influence decision-making and the development of sound terms and conditions

4.3.3 Mining sustainability and community relations

4.3.3.1 Mining sustainability


According to Wikipedia: Sustainability is the ability to exist constantly. In the 21st
century, it refers generally to the capacity for the biosphere and human civilization
to co-exist. It is also defined as the process of people maintaining change in a homeo-
stasis balanced environment, in which the exploitation of resources, the direction of
investments, the orientation of technological development and institutional change
are all in harmony and enhance both current and future potential to meet human
needs and aspirations. For many in the field, sustainability is defined through the
interconnected domains or pillars: environment, economic and social sustainability.
In the mining context, sustainability does not refer to the orebody itself as this will
always be depleted by mining; it rather refers to the long-term positive impact that the
mining operation should have mainly on the local community.
Historically, mines often focussed on quick and short-term projects and initiatives
rather than longer term and more sustainable projects (Ostensson & Roe, 2017). Good
corporate longer term programs typically include:

• Supply chain development


• Employment
• Local business development
• Regional development
• Infrastructure
Green and sustainable mining 361

Figure 4.7 Factors that influence mining sustainability

The International Council on Mining & Metals (ICMM) developed a Community


Development Toolkit (2005, updated 2012) (https://commdev.org/pdf/publications/
ICMM-Community-Development-Toolkit.pdf) whose main aim is:

By supporting communities to develop themselves in a sustainable manner,


a mining and metals company is simultaneously helping its own business to
succeed. Mining operations and their community development programs
should be viewed as a mutually beneficial partnership process to achieve
sustainability.

The aspects that impact mining sustainability is shown in Figure 4.7 (Mkandawire &
Oakes, 2015).

4.3.3.2 Community relations


According to the Mining, Minerals and Sustainable Development (MMSD) Project
facilitated by the International Institute for the Environment and Development
(2002): Mineral development can create new communities and bring wealth to those
already in existence, but it can also cause considerable disruption. New projects can
bring jobs, business activities, roads, schools and health clinics to remote and previ-
ously impoverished areas, but the benefits may be unevenly shared, and for some they
may be poor recompense for the loss of existing livelihoods and the damage to their
environment and culture. If communities feel they are being unfairly treated or inad-
equately compensated, mining can lead to social tension and sometimes to violent
conflict.
As the more accessible and higher-grade mineral deposits are mined and depleted,
exploration is moving to more remote areas and it is essential that these local
362 Green and sustainable mining

communities benefit in the short and long term from any mining projects. These
communities often are less developed and resourced so may need longer and more
in-depth engagement that must start with an understanding of their leadership struc-
ture, culture, values and aspirations.
The MMSD Project recommends the following for good community relations as a
minimum:

• A genuine commitment to sustainable development at the local level


• A commitment to effective community participation in decision-making
• A belief in open communication among actors
• A commitment to proactive rather than reactive approaches
• Respect for independent evaluation and monitoring systems
• A willingness to share responsibility and collaborate with others

Working out the boundaries of rights and responsibilities is the challenge; the precise
roles of the various participants will depend on local circumstances.

4.3.4 Social license to operate


It should be noted that a social license to operate is a concept rather than a legal
agreement or contract, but it is important because it shows that the stakeholders,
particularly the local community, support the specific proposed or active mining
operation.
According to Tyson (2018), corporate social responsibility (CSR) was developed in
the 1930s and focussed on how mines behaved ethically, legally, economically and phil-
anthropically. From about the 1980s, however, local communities started to demand
more and become more directly involved especially as far as mining-induced environ-
mental impact was concerned. This was helped by the internet that made information
much more available to all. The concept of mine owners deciding what communities
would be offered and when was not acceptable and that a social partnership that was
transparent was essential. Figure 4.8 shows the value of mines spend on CSR. This
was the result of local communities and more broadly society in general becoming
more aware and educated on the need to protect the environment.
Mining companies found that not involving the local communities in a mean-
ingful dialogue that resulted in concrete actions caused increased costs from both
legal and “less legal” community actions that often-included blockades and active
protects.
It is generally accepted that James Cooney, a Director of Canadian-based Placer
Dome, coined the concept of “Social License to Operate” in 1997 after a major tailings
collapse on one of its mines in the Philippines in 1996.
Joyce and Thomson (2000) stated that:

A social license to operate exists when a mineral exploration or mining project


is seen as having the approval, the broad acceptance of society to conduct its
activities … Such acceptability must be achieved on many levels, but it must begin
with, and be firmly grounded in, the social acceptance of the resource develop-
ment by local communities.
Green and sustainable mining 363

Figure 4.8 The value of corporate spending on corporate social responsibility


Source: Tyson, 2018

Mining operations therefore now need approvals not only from the relevant gov-
ernment of the day but also from the local community (and society in general to some
extent). The government approval involves legal permitting and formal regulation,
whereas the social license with the local community has no legal standing but is prob-
ably as important and requires good community relations and openness to ensure
benefits for all stakeholders.
According to Cooney (2016), the “Three markers of social licence are”:

• An ongoing structured process of consultation between a mining company and


the local communities, together with their allies
• A “precautionary approach” (beyond regulatory compliance) to deal with low
probability but potentially high-impact environmental disasters
• A mutually agreed plan to optimize the economic benefits from a mine to local
communities within a long-term future vision

He further states that it may involve one or more of the following:

• Corporate social responsibility


• Sustainable economic, environmental and social development
• Community rights and entitlements
• Social justice: distributional and procedural fairness
• Evolution in the decision-making power of government
• A new social contract that legitimizes corporations by redefining their obligations
to society
• Any broad public policy issue that is not addressed in government approval pro-
cesses for industrial projects
364 Green and sustainable mining

The International Standard Organization (2010) defined it as:

Responsibility of an organization for the impacts of its decisions and activities


on society and the environment, through transparent and ethical behaviour that:

• Contributes to sustainable development


• Takes into account the expectations of stakeholders
• Is in compliance with applicable law and consistent with international norms
of behaviour
• Is integrated throughout the organization and practised in its relations

Mines must now not just comply with laws and regulations but also with ethically
acceptable norms as defined by society and public perceptions (which rightly or
wrongly equates to their reality).

4.4 Mine waste disposal and repurposing

4.4.1 Waste dam design and disposal

4.4.1.1 Introduction
Almost all mines, whether surface or underground, produce a finely ground waste
material called tailings, after the mineral or metal of interest has been removed during
the metallurgical process. These waste materials can be disposed of in the following
ways:

• Into the sea or a lake (this method is not now favoured due to environmental
pressures)
• Into a river behind a suitable dam wall (this can be a costly method and also has
adverse environmental considerations)
• Onto a specifically constructed tailings dam on surface
• Placing the waste back underground as a backfill material (as mentioned later,
this method cannot always be used to remove all the tailings produced, and some
surface disposal is still required)

At the beginning of 2020, the total volume of tailings stored in facilities disclosed by
mining companies representing 54% of the global mining industry by market capitali-
zation was 45.7 billion m³, according to a study by Prof. Elaine Baker of the University
of Sydney and GRID-Arendal, a non-profit environmental communications centre
based in Norway. This figure was projected to reach more than 56.6 billion m³ by 2025,
a 25% increase.
Figure 4.9 shows a typical large hard rock tailings disposal dam.
TYPICAL GOLD, PLATINUM AND CHROME TAILINGS DISPOSAL FACILITIES

If the topography does not allow for disposal in a valley-type structure, then a con-
ventional tailings dam is required. In order to accommodate the tailings, containment
paddocks formed by earth bunds or thickened and coarser (classified) tailings are con-
structed into which the tailings is pumped from the processing plant. These paddocks
Green and sustainable mining 365

Figure 4.9 A platinum tailings dam, Marula Mine


Reproduced with kind permission by Impala Platinum Holdings Ltd.

act as sedimentation/settlement ponds from which water can be ideally recovered for
reuse, discharged (after treatment) or allowed to evaporate.
As time progresses and the initial paddocks become full, their height is extended by
packing out the outer perimeter often using the previously deposited tailings mate-
rial. Further advancements to this simple method of disposal were made in order to
minimize the costs associated with progressively raising the outer perimeter. This was
achieved by providing multiple tailings discharge points around the perimeter and the
construction of a perimeter paddock. The deposition of tailings was then controlled
such that the perimeter paddock was maintained 1.0–1.5m above the level of the set-
tlement basin. This was achieved by preferential deposition of the coarse fraction of
the tailings stream into the outer paddock and allowing the water containing the fine
fraction to flow through the paddock and onto the central basin, where water could
be recovered. In order to adequately control this type of deposition, deposition into
the perimeter paddock was limited to the daylight hours with night-time deposition
occurring within the basin. This necessitated paired discharge pipes at each discharge
point, one to serve the paddock (or “day-wall”) and one to serve the basin. The decant
water from the slurry tailings deposition and any storm water accumulated on the
impoundment for return to the plant for reuse was achieved via a central gravity pen-
stock. This method of impoundment construction is still widely used in South African
gold mines; however, more modern designs include various drainage and monitoring
systems to ensure impoundment stability. The construction of the day-wall/basin sys-
tem is necessary as it provides adequate freeboard to accommodate huge infrequent
rain events that would otherwise not be achievable due to the near-flat beach angle of
the mine tailings. Changes in weather created by global warming means that more
freeboard should be provided to safely handle these huge downpours of rain, which in
many areas are becoming more common.
366 Green and sustainable mining

TYPICAL DIAMOND TAILINGS DISPOSAL FACILITIES

The fine tailings generated through crushing and washing of kimberlite ores is not
suitable for disposal by conventional methods. This is due to the very fine nature and
poor sedimentation and drainage characteristics. Typically, diamond tailings is dis-
posed of in a pre-constructed impoundment formed by compacted earthfill or within
worked-out pits. Open-end discharge from the tailings delivery column is common
with barge mounted pumps required for the return of process and stormwater.
TYPICAL COAL, IRON ORE AND BASE METALS TAILINGS DISPOSAL FACILITIES

The disposal of each of these materials is dependent on the specific material charac-
teristics encountered in each mine. Typically, disposal takes the form of either spigot
deposition, open-end deposition or a slightly modified version of either. These modifi-
cations frequently incorporate cyclone deposition as a method of perimeter embank-
ment construction. Under certain circumstances, cyclone deposition has been utilized
for tailings disposal.
All hard rock metal mines produce much more tailings than can be disposed of as
backfilling because of the very low percentage of metal recovery and the bulking of
the blasted ore (around 30%).
In regions where water is short, tailings dams put an additional severe strain onto
the water availability. According to Matthews and Spearing (1999), tailings disposal
constitutes the single longest consumer of water in South African mines. Typical fig-
ures for the platinum industry are 65% of all water consumed (i.e., the water lost from
the mine water circuit) is attributable to tailings disposal, and of the process water
pumped to the tailings dam, 35–40% is recovered for reuse.

4.4.1.2 Engineering considerations for tailings dam design


There are many considerations to be taken into account when planning a tailings
basin especially with a new mining operation or in the case of a major expansion to an
existing operation. Some of the considerations for planning a tailings dam and basin
are:

• The life and size of the facility


• Water reuse needs
• Environmental issues
• Monitoring requirements
• Operational and functional safety
• Land availability and topography
• Cost of construction and operation
• Future reclamation

Some of the considerations for dam design once the location is finalized are:

• Spillway location, type and sizes


• Water reclamation methods
• Availability of materials for dam construction
• Controlling seepage (dam failure and pollution from leakage)
Green and sustainable mining 367

• Foundations
• Abutment
• Operations

4.4.1.3 Conventional dam construction methods


There are several methods of tailings dam construction that are used. The selected
method is a function of:

• Terrain
• Operations
• Availability of materials
• Cost

The common methods include:

• Conventional method
• Downstream method
• Upstream method
• Centreline method
CONVENTIONAL METHOD

In this method, a conventional earth dam wall is constructed in a suitable valley or


ravine and the tailings stored behind the single wall.
DOWNSTREAM METHOD

As the volume of tailings stored increases with time, the retaining wall height is
increased and the centre of gravity of the wall moves away from the centre of the dam
(i.e. down dip).
According to Engels and Dixon-Hardy (2007), the downstream design was devel-
oped to reduce the risks associated with the upstream design, particularly when
subjected to dynamic loading as a result of earthquake shaking. The installation of
impervious cores and drainage zones can also allow the impoundment to hold a sub-
stantial volume of water directly against the upstream face of the embankment with-
out jeopardizing stability.
The downstream embankment design starts with an impervious starter dyke
unlike the upstream design that has a pervious starter dyke. The tailings are at
first deposited behind the dyke, and as the embankment is raised, the new wall is
constructed and supported on top of the downstream slope of the previous sec-
tion (Figure 4.10). This shifts the centreline of the top of the dam downstream as
the embankment stages are progressively raised (Vick, 1990). An advantage of the
downstream design is that the raised sections can be designed to be of variable
porosity to tackle any problems with the phreatic surface of the embankment. This
can be particularly useful where a processing plant has made changes to increase
efficiency and as a result the tailings characteristics are altered. This may result in
pumping more water to the tailings facility or alter the drainage characteristics of
the newly deposited tailings.
368 Green and sustainable mining

Figure 4.10 Downstream method (Engels & Dixon-Hardy, 2007)

UPSTREAM METHOD

This is the opposite of the downstream method as the centreline of the wall moves into
the dam as the wall height increases.
According to Engels and Dixon-Hardy (2007), the upstream method is the lowest
initial cost and most popular design for a raised tailings embankment. One of the rea-
sons for this is mainly due to the minimal amount of fill material required for initial
construction and subsequent raising, which normally consists entirely of the coarse
fraction of the tailings (created by classification).
The construction of an upstream designed embankment starts with a pervious (free
draining) starter dyke foundation. The tailings are usually discharged from the top of
the dam crest, creating a beach that becomes the foundation for future embankment
raises (Vick, 1990). Figure 4.11 shows a simplistic diagram of the stages of construction
of an upstream raised embankment. Where the tailings properties are suitable, natu-
ral segregation of coarse material settles closest to the spigot (near the wall) and the
finest material furthest away (near the dam centre). Cyclones can be used to accelerate
this particle segregation for certain tailings characteristics to send the slime propor-
tion to the centre of the impoundment and the sand fraction to the beach behind the
crest. The conventional method of upstream lifts relies on no compaction of the spig-
otted beach that forms the embankment shell (Martin & McRoberts, 1999). Today,
compaction by earthmoving equipment is common to increase the degree of safety of
Green and sustainable mining 369

Figure 4.11 Upstream method (Engels & Dixon-Hardy, 2007)

raised embankments. Generally, the settled coarse fraction from the spigots/discharge
points is used as the raise material for the embankments. For multiple spigot discharge,
a series of shallow pits are dug in front of the spigots (once the tailings have dried and
consolidated) and tailings placed on the embankment crest, which are then compacted,
the tailings lines lifted and reassembled and then normal operation commences.
It is not surprising that the upstream method is the most common design to fail, caus-
ing huge environmental consequences all over the world (International Commission on
Large Dams (ICOLD) and United Nations Environment Programme (UNEP), 2021).
CENTRELINE METHOD

The centreline method is a compromise between both the upstream and downstream
designs (Benckert & Eurenius, 2001). It is more stable than the upstream method and
does not require as much construction material as the downstream design. Like the
upstream method, the tailings are generally discharged by spigots from the embank-
ment crest to form a beach behind the dam wall. When subsequent raising is required,
material is placed on both the tailings and the existing embankment. The embankment
crest is being raised vertically and does not move in relation to the upstream and down-
stream directions of subsequent raises, hence the term centreline design (Figure 4.12).
The design incorporates the internal drainage zones that are similarly found in
the downstream method. Therefore, the free water can encroach closer to the dam
crest than in the upstream method without the worry of raising the phreatic surface
370 Green and sustainable mining

Figure 4.12 Centreline method (Engels & Dixon-Hardy, 2007)

and causing a potential failure risk. However, a centreline dam cannot be used as a
large water retention facility solely due to the subsequent raises being partially built
on consolidated tailings. A suitable decant system needs to be installed to prevent
the free water submerging the beach around the dam crest, which is hazardous.

4.4.2 Waste dam operation and monitoring

4.4.2.1 Dam operation


The tailings dam operators often control the tailings dam design, with respect to the
method of dyke construction and availability of materials.
Some of the considerations for tailings dam operations are:

• Water pond location and elevation (within the dam)


• Spillway
• Tailings deposition (placement)

The problems with conventional tailings dams include the following:

• The inherent safety risk of mainly liquefaction, such as what occurred at a Vale-owned
iron ore mine in Brazil in 2019, causing severe environmental damage and loss of life
Green and sustainable mining 371

• Serious water losses on the dams in regions where water is a scarce commodity
(such as in Australia, Northern China, the Middle East and South Africa)
• The large area needed for the dam and the cost of eventual reclamation

Many mines are already considering high-density (paste) disposal for the above rea-
sons. The international mining industry is so concerned with current surface tailings
disposal problems that several international conferences have already been held in
places such as Canada, Australia and South Africa.
Figure 4.13 (Robinsky, 1999) compares the conventional surface tailings method
with the high-density method (in which foam technology could be a key element to
replace some of the transport water traditionally used). Figure 4.14 shows the sche-
matic of a conventional tailings dam.

Figure 4.13 Comparison of conventional tailings disposal and tailings dry stacking
372 Green and sustainable mining

Figure 4.14 A conventional tailings dam

4.4.2.2 Main causes for tailings dam failures


The main causes of dam failures are:
• Overtopping of the walls mainly caused by a huge rain event
• Dam wall failure
• Structural failure often caused by seismicity
• Surface erosion over time caused by weathering
• Piping erosion (a similar dangerous phenomenon when backfilling large stopes)
• Design flaws caused by unknown or unplanned conditions
Any monitoring program therefore needs to address the main causes of dam failure to
warn of potential problems in advance of a failure. Obviously avoiding any failure is
the prime goal but effective monitoring is also essential in all cases.

4.4.2.3 Tailings dam monitoring


Unlike most water dams, tailings dams are not static and typically grow until they
reach some predetermined maximum “safe height”. Monitoring programs and safety
Green and sustainable mining 373

audits therefore need to be undertaken regularly during the active life of the tailings
dam and for some time afterwards until it has been determined to be “safe” and no
longer poses an environmental hazard.
Most of the monitoring is done visually (looking for leaks, etc.) using piezometers
(to monitor pore water pressure changes) and more recently radar or other technolo-
gies able to detect very small movements.
BHP (2019) uses the following to monitor and review the stability of dams they
operate:

• Routine surveillance by the specific operators at the mine on a daily or weekly basis
• Dam safety inspections by the responsible dam engineer on a weekly or longer
time frame depending on the specific dam
• Dam safety inspections by the external engineer of record who reviews all the
internal inspections, analyses the monitoring data and reviews all the relevant
policies and procedures annually
• Dam safety reviews by an external professional engineer who every 5–10 years
checks the dam’s integrity against the design, construction, performance, opera-
tion and emergency planning procedures
• Tailings Review Boards that consist of independent experts who meet as needed
depending on any inherent risks

The use of external independent experts in the monitoring and reviews seems to have
great merit.

4.4.3 Tailings dam safety, design and operation in the future

4.4.3.1 Tailings major hazards


A tailings dam failure is one of the largest disasters that can occur on a mine and can
cause potentially massive loss of life, environmental damage and the loss of public
confidence in mining. Unfortunately, there have been regular failures globally, but the
January 2019 failure of the Vale-owned Brumadinho iron ore dam disaster in Brazil
was a final serious call to action by mines globally. This collapse killed 270 people and
released over 10 million m³ of waste, and followed the November 2015 failure of Vale
and BHP’s joint venture (JV) Fundao tailings dam that killed 19 people also in Brazil
(Owen et al., 2020).
These catastrophic failures caused the International Council on Mining & Metals
ICMM to create a global taskforce with other partners, including the UNEP, to inves-
tigate the problem and propose solutions. This resulted in the August 2020 report
titled the “Global Industry Standard on Tailings Management”.
This Global Industry Standard mandates that mine operators take responsibility
for the safety of their tailings storage facilities (TSFs) through all phases of a facil-
ity’s lifecycle – this should have always been the status quo. It includes closure and
post-closure so that the goal of zero harm to people and the environment can hope-
fully be achieved.
The majority of the dam failures have been linked to the upstream dam wall raising
method. According to GRID-Arendal (2020), this method has accounted for 43% of
historically recorded dam collapses.
374 Green and sustainable mining

4.4.3.2 Tailings dam future operations


Following the devastating failure in Brazil in 2019, many sound recommendations
have been made to try and ensure that such massive disasters don’t recur in the future.
According to Brereton et al. (2020), tailings dam planning in the future must include
the following elements:

• Human rights issues associated with the local communities


• Meaningful community engagement throughout the planning, operation and
closure of a tailings dam
• Detailed EIAs focussing on the proposed tailings dam
• Baseline studies
• Human exposure investigations
• Social considerations associated with any involuntary resettlements that may be
needed
• Emergency planning
• Long-term recovery after the tailings dam’s operation stops

In addition, the authors believe that the planning of a new tailings dam should also
include:

• Global best practices investigation


• Consideration of selective mining and sorting before comminution to reduce tail-
ings production
• Partial or total alternative use for the tailings as backfill for example
• Any future repurposing options

Oberle (2020) from the Global Tailings Review lists the following Principles that dam
operators shall comply with going forward:

• Principle 1: Respect the rights of project affected people and meaningfully engage
with them at all phases of the tailing’s facility lifecycle including closure.
• Principle 2: Develop and maintain an interdisciplinary knowledge base to sup-
port safe tailings management throughout the tailing’s facility lifecycle, including
closure.
• Principle 3: Use all elements of the knowledge base – social, environmental, local
economic and technical – to inform decisions throughout the tailing’s facility life-
cycle, including closure.
• Principle 4: Develop plans and design criteria for the tailings facility to minimise
risk for all phases of its lifecycle, including closure and post closure.
• Principle 5: Develop a robust design that integrates the knowledge base and min-
imises the risk of failure to people and the environment for all phases of the tail-
ing’s facility lifecycle, including closure and post closure.
• Principle 6: Plan, build and operate the tailings facility to manage risk at all
phases of the tailing’s facility lifecycle, including closure and post closure.
• Principle 7: Design, implement and operate monitoring systems to manage risk at
all phases of the facility lifecycle, including closure.
• Principle 8: Establish policies, systems and accountabilities to support the safety
and integrity of the tailings facility.
Green and sustainable mining 375

• Principle 9: Appoint and empower an engineer of record.


• Principle 10: Establish and implement levels of review as part of a strong quality
and risk management system for all phases of the tailing’s facility lifecycle, includ-
ing closure.
• Principle 11: Develop an organisational culture that promotes learning, commu-
nication and early problem recognition.
• Principle 12: Establish a process for reporting and addressing concerns and imple-
ment whistle-blower protections.
• Principle 13: Prepare for emergency response to tailings facility failures.
• Principle 14: Prepare for long term recovery in the event of catastrophic failure.
• Principle 15: Publicly disclose and provide access to information about the tail-
ings facility to support public accountability.

Densification of the tailings prior to deposition on a dam seems necessary in the


future even though it increases the disposal costs. Figure 4.15 shows the situation fre-
quently used at present, and Figures 4.16 and 4.17 represent options where the place-
ment density is increased, which improves the stability of the whole dam structure
(Vietti, 2020).
It is likely that tailings disposal facilities in the future will involve dense (paste)
type disposal, which appears less of a failure risk and easier to effectively rehabilitate,
although more costly.

Figure 4.15 The current typical tailings disposal schematic (Vietti, 2020)

Figure 4.16 High-density tailings disposal (Vietti, 2020)


376 Green and sustainable mining

Figure 4.17 Two-stage higher density tailings disposal (Vietti, 2020)

4.4.4 Waste dam rehabilitation


The planning for the design and operation of a tailings basin should consider the
reclamation of the deposition of tailings, location of spillways and final drainage
patterns.
Rehabilitation measures usually include:

• Recontouring if safe to do so into a more useful topography.


• Stabilizing the outer walls.
• Establishing runoff water collection facilities so the quality can be made accept-
able before it is released.
• Improving the wall and surface ability to sustain vegetation. This often involves
the placement of soil, fertilizers and drip irrigation to encourage growth of gener-
ally very hardy grasses at least initially.

In recent years, the costs associated with dam rehabilitation have increased dra-
matically due mainly to the environmental lobby in first-world and developed
countries. A single large tailings dam at the now defunct Elliot Lake Mine cost
over US$100 million to rehabilitate at the end of the life of the mine (Nicholson
et al., 2012).

4.4.5 Repurposing options and utilization


Generally, tailings dams contain fine to very fine particles of the materials within the
orebody after most of the desirable metal or mineral has been extracted. Metal mines
typically produce the most tailings and consist mainly of fine materials that could be
used in construction as aggregates and often iron oxides.
If the metal mine tailings dam is relatively old, new process technology may also
liberate additional metals of value that can at least offset some or all of the repurpos-
ing costs. These could include the original target metal mined and also possibly rare
earths and lithium as examples. These opportunities, however, do not reduce the vol-
ume of the tailings and only by producing fine aggregates, useful clays and iron oxide
can the tailings dam be effectively eliminated.
Green and sustainable mining 377

4.4.6 Challenges to repurpose tailings cost effectively


The three biggest challenges associated with repurposing tailings dams are:

• The reprocessing of material from an existing TSF in itself can increase the likeli-
hood of a failure so the design must be detailed and comprehensive, involving all
stakeholders (especially the local community).
• Location, as generally the mines are remote from where the repurposed materials
could be used, making it uneconomic from a transportation cost point of view.
• Having too much of a saleable repurposed product, resulting in little of the dam being
actually reprocessed, and this could increase the potential for a failure, depending on
how and from where on the dam the material for repurposing was taken.

4.5 Urban mining

4.5.1 Introduction to urban mining


Urban mining can be defined as reprocessing urban waste to create materials of value
for reuse.
As shallow and higher grade orebodies become exhausted, mining becomes deeper
and typically at lower grades. This increases the costs and safety hazards and so makes
the incentives to start urban mining much higher. It is also a potential advantage for
natural resource-poor countries to obtain some of their raw materials internally and
therefore reduce imports.
Waste is also a critical issue in general but especially in the metals value chain, from
mining waste to eventual end-of-life products such as scrap steel from construction
and demolition waste or the growing problem of electronic waste. The goal is the cre-
ation of “zero waste” ultimately using a true circular economy.
Urban mining is a relatively new concept and is the process of reclaiming useful
compounds, materials and elements from products, buildings and waste rather than
disposing them off after their typical useful life span. The value chain for urban min-
ing is shown in Figure 4.18. It is an important part of sustainable development. Whilst
the amount of urban waste is a serious challenge, it can become an opportunity too.
Urban mining can be undertaken in two different ways:

• From existing landfills or waste dumps


• Collecting products and waste at the source (e.g., a household, company or factory)

The collection tends to be the issue with most recycling.

Figure 4.18 The value chain for urban mining


378 Green and sustainable mining

Figure 4.19 Raw material supply in a circular mining economy (Medkova, 2016)

Figure 4.19 shows how conventional mining and urban mining mutually contribute
to the raw material supply chain, and Figure 4.20 shows the outline of a comprehen-
sive urban mining system.

4.5.2 Waste suitability for urban mining


Wastes that are suitable for urban mining include:

• Plastics, paper and glass


• Building and construction debris
• Electronic waste

Figure 4.21 shows the potential for a metal-based circular mining economy.

4.5.2.1 Plastics, paper and glass


PLASTICS (POLYMERS)

Plastics (polymers) are organic materials made from coal, oils or natural gas which are
all non-renewable natural resources.
According to Tiseo (2020), the plastics industry began in the early 1900s when the
first synthetic plastic was created by Leo Hendrik Baekeland in the US. Since the
industry began, annual production has exploded from some 1.5 million metric tonnes
in 1950 to 359 billion metric tonnes in 2018. The cumulative production of plastic has
Green and sustainable mining 379

Figure 4.20 Urban mining outline


Source: Urban Mining Organization

already surpassed 8 billion metric tonnes worldwide, with further increases expected
in the coming decades.
Recycling plastics has obvious advantages, including:

• Less environmental pollution (especially of waterways and oceans)


• Being more cost effective than producing new plastic materials
• Extending the useful life of scarce raw materials to make plastic (mainly oil)
• Provides a more sustainable resource

All plastics can be recycled but not all economically at present. The most commonly
recycled plastics are polyethylene terephthalate (PET) and high-density polyethylene
(HDPE), which are used mainly to make bottles.
380 Green and sustainable mining

Figure 4.21 The metal circular economy


Source: Material Trader

Much of the plastic waste finds its way into the oceans’ eco-systems as tiny plastic
pieces where it is causing immense environmental damage (Figure 4.22).
The amount of plastics reaching waterways and the oceans is increasing as shown
in Figure 4.23.
Plastic recycling is growing but more needs to be done quickly and perhaps more
regulation is required to coerce users, manufacturers and merchants to utilize more
biodegradable options.
The main issue with plastic (as with most products for recycling) is the collection
and transport to a sorting and recycling facility. It is then typically cleaned, heated and
formed into resin pellets that can be reused. The pellets can then be used to make new
plastic products by themselves or blended with new plastic pellets. Chemical reprocess-
ing is also becoming more common to reprocess the plastic waste. Figure 4.24 shows
global recycling statistics (Frerot, 2018).
PAPER

Paper is derived from wood fibre and requires chemical additives to liberate the lignin.
It can be considered a renewable material if the timber is obtained from commercial
forests.
Paper recycling is common, but because the fibre length reduces during recycling, it
is generally blended with new fibre to produce a new product. After several recycling
Green and sustainable mining 381

Figure 4.22 Plastic pollution in the water


Reproduced with kind permission from Foundation for Information & Research on Marine Mammals

Figure 4.23 Coastal plastic waste produced


Source: Climate Desk
382 Green and sustainable mining

Figure 4.24 Selected global plastic recycling statistics (Frerot, 2018)

processes, the fibres are, however, generally too short. Since paper is carbon based,
the more it can be recycled, the better for the environment. Tissue, towels and waxed
papers are generally not recycled.
GLASS

Glass is made from sand (silica), lime and soda (sodium carbonate), which when
heated together form a liquid which can be moulded or formed into many glass items.
Glass is totally recyclable, and the process is simple. The glass is washed, separated
by colour and type, then crushed and heated into a liquid again. Removing impurities
is the only challenge, and this is an important part of recycling.

4.5.2.2 Building and construction debris


When buildings are demolished for whatever reason, the potential to recycle most of
the materials is great. This includes:

• Concrete that can be crushed, reprocessed and reused


• Asphalt that can be crushed, heated and the asphalt reused
• Metals including steel, copper, tin, lead and zinc.

The building can be demolished using blasting or other means and the metal (mostly
steel used for concrete reinforcement) is removed typically using a magnet or by hand
(especially for non-magnetic metals).

4.5.2.3 Electronic waste


The demand for electronics and the technology-driven products is increasing very
quickly. This means that the demand for “new and more powerful” products is increas-
ing, the supply of new raw materials is under pressure, the product life is reducing and
the waste generated is becoming a very serious issue. The average life of a mobile
Green and sustainable mining 383

phone is now typically only 2–3 years. Gold, platinum group metals and rare earths
can be “mined” from electronic waste that includes desktop and laptop computers and
mobile phones.
It would help if electronic manufacturers built more modular equipment such
that updates could be incorporated into older models without the need to totally
replace the older model. This would, however, adversely affect the profitability of
the companies but the environmental advantages would be huge. Standardization
of basic components such as frames, batteries, plugs and chargers across manufac-
turers would be very helpful too, although again not favoured by the companies for
obvious reasons.
It is interesting to note that the gold in mobile phones is typically at a grade 100 times
that which is typically conventionally mined. From a mobile phone, gold, nickel, alu-
minium, steel and rare earths can be recovered economically if the collection system
is set up effectively.
According to Rau (2019), a million mobile phones when reprocessed will yield:

• 16,000 kg of copper
• 350 kg of silver
• 35 kg of gold
• 15 kg of palladium

A recent success story is UK-based Advanced Plasma Power Ltd (APP), which spe-
cializes in turning waste to energy and fuels. They have partnered to recycle a giant
landfill in Belgium. With local partners, they will mine a giant landfill containing
considerable e-waste that is decades old. They will produce two economic streams:
recycled materials and methane gas to be used for power generation. Hopefully this
business model will be replicated elsewhere to not only recycle waste, but to produce
a new sustainable industry that will be vital in developing a viable circular economy.

4.5.2.4 Battery waste


The safe recycling of spent non-rechargeable batteries is important as the alternative
is serious environment damage. In 2009, Europe banned nickel-cadmium batteries
due to the toxicity of especially cadmium. Most lead-based automotive batteries are
recycled in developed countries as about 70% of the weight is lead, which can be easily
recycled. Not enough mobile phone batteries are recycled mainly due to current cost
and collection considerations.
Once batteries are collected, they need to be sorted at a central facility based on
their chemistry, broadly into:

• Lead
• Nickel-cadmium
• Nickel-metal hybrid
• Primary lithium and lithium ion

Sadly, with the exception of lead batteries, other batteries take more energy to repro-
cess currently than to produce new ones.
384 Green and sustainable mining

4.5.3 Challenges to growing urban mining


Urban mining makes sense for the environment as the current standard method of
sending “e-waste” to a waste fill is not only wasteful but harms the environment by
polluting the ground water with acids and heavy metals. In addition, all recycling or
reprocessing reduces the amount of finite new materials that need to be mined and
generally requires much less energy to reprocess.
The biggest challenge to all recycling efforts is getting the typically small quantities
of waste materials from a huge number of geographical sites to a central processing
facility with the necessary economies of scale.
According to a blog by Meilani (2018):

There are a multitude of regulations that businesses and regions must comply to,
and for good reason. These regulations keep our waste treated and our products
up to a standard. Nonetheless, these same regulations may be the ones standing
in the way of a circular economy, intentionally or not. This is due firstly to the
sheer amount of regulations that govern the industrial and economic world. This
means companies and industries wishing to work towards a circular economy
are trapped by not one but multiple various policies that in some way restrict or
counteract efforts towards going circular. And even if one regulation undergoes a
change for the better, they likely still find themselves blocked by the effects of all
the others.
However, it’s not as if all these limiting regulations purposely and actively seek
to inhibit circular activities – many times the blockade is simply a side-effect of
the regulation. For example, certain regulations induce waste, especially in the
food sector We should still push policy-makers to make tweaks and amendments
that would enable the regulations to allow for the achievement of both their pri-
mary purposes and a circular economy.

Incentives need to be offered to consumers to correctly handle their old electronics


and send them for reprocessing rather than to the waste dump.
A further problem is that waste typically contains a mixture of materials, both
inorganic and organic, and that collection and transportation to reprocessing
facilities can be complex and costly depending on many factors. These include
the local population density, the infrastructure and local economy. Waste clas-
sification is important and requires complex sorting and can consist of (often a
combination of):

• Domestic waste
• Industrial waste (that is often toxic)
• Medical waste
• Radioactive waste
• Agricultural waste
• Electronic waste
• Organic waste

Reworking old waste and landfills also pose potentially high risks for the workforce
involved due to, for example, dust, toxic leaching and gas build-ups.
Green and sustainable mining 385

4.6 Mining risk and risk management

4.6.1 Introduction to Risk


According to Rhodes (2020):

Some historians believe that the earliest concept of managing risk arose because of
gaming. Thousands of years before Internet users could play online poker, people in
different ancient civilizations played games with dice and bones. These games evolved
into more complex games such as chess and checkers well over 2,000 years ago. Some
historical evidence that gaming gave rise to probability theory, important to risk
management, comes from writings by Dante (in the 1300s) and Galileo (1564 to 1642).
The famous mathematicians, Pascal and Fermat, wrote each other about games of
chance in the 1600s, a correspondence that is believed to have given rise to the mod-
ern probability theory used today. If one were to consider the role of insurance in risk
management, it is still possible to trace it back to ancient times. For example, mutual
aid and burial societies have been documented as far back as the earliest days of
ancient Rome. These are considered the precursors of modern insurance companies.

The history of modern risk is therefore as follows:

• The modern concept used in mining was basically developed by the insurance and
financial institutions so they could ensure a profitable insurance-based business model.
• The insurance coverage offered and the associated risks had to be quantified and
understood in order to still make an overall profit.

It should be noted that risk assessments have also been used in the prison system for dec-
ades to try and assess if a prisoner could be “safely released” and not re-offend if paroled.
A “risk free” environment does not exist, so the objective therefore is to understand
and quantify the risk and eliminate or mitigate it.
The reasons to effectively manage risk include:

• To prevent accidents
• To reduce medical costs
• To reduce workman’s compensation premiums
• To reduce insurance premiums
• Keep employee morale up
• To maintain a positive image with the community and the media
• Accidents, incidents, injuries, medical treatment, loss of experienced personnel
and a negative image with the media and community cost the company financially
and can damage its reputation.

Generally, risk is related to the expected losses which can be caused by a risky event
and to the probability of this event:

• The harsher the loss and the more likely the event, the worse the risk.
• Measuring risk is often difficult; rare failures can be hard to estimate, and loss of
human life is generally considered irreplaceable.
386 Green and sustainable mining

Figure 4.25 The risk matrix

The engineering definition of risk is:

Risk = (probability of an accident) × (losses per accident)

Figure 4.25 shows the risk matrix which prioritises the risk into areas. A key in any
industry is that after an incident or accident, it is vital that a recurrence is avoided. A
detailed investigation to identify the source and root cause of the incident or accident
is important.

4.6.2 Mining hazards


Every industry has risk associated with it and the mining industry is no exception.
Managing the risk is particularly important in mining because the “material” that
is worked in is not an engineered product; it is natural and hence highly variable, and
often the localized conditions are unknown. In addition, mining is:

• Inherently dangerous, especially underground where the workforce works under


a “rock” roof.
• Multiple activities take place at the same time and often in cramped conditions.
• Employees are regularly in close contact with large mobile and fixed equipment.
• Equipment and processes are expensive and complex and need to be coordinated.
• Unplanned events occur routinely, and the impact of these unplanned events can
be significant.
• The probability of loss is high.
• The cost of an incident, accident or loss can be high.

4.6.2.1 Surface mining hazards


Some of the hazards would include:

• Dust exposure
• Noise exposure
• Ultraviolet (UV) and thermal exposure
• Long-term injury from ergonomic-type issues (e.g., sitting driving equipment for
extended time periods)
Green and sustainable mining 387

• Exposure to hazardous minerals, fuels and chemicals


• Highwall instability
• Blasting dangers (especially from fly-rock)
• Material handling issues
• Electrical dangers
• Effects of adverse weather (especially floods and thunderstorms)
• Fatigue
• Slips and falls
• Security issues (especially keeping the public off-site)
• Waste rock dumps and tailings dam stability
• Water contamination

4.6.2.2 Underground mining hazards


Basically, all of the hazards associated with surface mining are common with under-
ground mines, but underground mines have additional potential hazards, including:

• Falls of ground
• Seismicity
• Water inrush
• Fire
• Surface damage and subsidence
• Mine access security
• Working in confined spaces
• Communication and visibility challenges
• Ventilation issues (especially those associated with poisonous/noxious or explosive
gas build-up)

4.6.2.3 Tolerable risk


Tolerable risk is defined differently by every business or industry and varies country
by country. Every business in operation today utilizes some type of risk management
program.
Tolerable risk depends on:

• Local social norms


• The particular occupation
• Industry standards
• Federal, local and government regulations
• Practices of industry partners
• Qualitative assessment of what is fair and reasonable
• Insurance mandates
• The comfort level for the company officials
• The history in the country and the industry

Acceptable risk for certain occupations such as the military, police, fire, emergency medi-
cal personnel, prison staff, rescue members, medical doctors and nurses are much higher
than for most other occupations due to the demands and requirements of their jobs.
388 Green and sustainable mining

In third-world and developing countries, acceptable risks is generally much higher


(i.e., more hazardous) than in first-world or developed countries. This is because many
are un-employed or under-employed and just battle to survive, so any job is acceptable
irrespective of the associated risk.
Effective training is a vital component in managing and reducing the risks and haz-
ards associated with both surface and underground mining. Management must also
be totally committed to safety through their actions and the target must always be
“zero harm” as the goal.
Safety reduction and risk management should be the key in and out of the work-
place. It is interesting to note that most workers do not follow their safety training at
home after work. An example is that at work everyone must use three points of contact
when climbing up and going downstairs. Very few do that outside the workplace yet
the risk remains the same!

4.6.2.4 Overcoming the fear of change


People seldom like or readily accept change and usually resist it. Risk management
often involves bringing in new rules, policies and procedures and it requires a mindset
check.
The following can help getting the vital workforce buy-in needed to improve safety
by reducing risk:

• Create an inclusive and effective safety committee


• Actively request and discuss all employee’s concerns to try and avoid resistance
• Actively involve all employees
• Listen to employees’ ideas
• Let the employees help make the final safety-related decision

Management must also overcome their hesitation to hand over at least partial control
of risk and safety and share it with the workforce.

4.6.2.5 Identifying risks


Risks can be identified using prior experience and data sources such as:

• Workman’s Compensation Records (or the equivalent in any country)


• Medical reportables and accidents
• First-aid cases
• Industry records
• Records of unplanned incidents and events

4.6.3 Risk management


Merriam Webster’s definition of “risk” is:

• Possibility of loss or injury: peril


• Someone or something that creates or suggests a hazard
Green and sustainable mining 389

• The chance of loss or the perils to the subject matter of an insurance contract,
and also:
• The degree of probability of such loss
• A person or thing that is a specified hazard to an insurer
• An insurance hazard from a specified cause or source (war risk)
• The chance that an investment (such as a stock or commodity) will lose
money
Merriam Webster’s definition of “management” is:

• The act or skill of controlling and making decisions about a business, department,
sports team, etc.
• The people who make decisions about a business, department, sports team, etc.
• The act or process of deciding how to use something

Generally, risk management is the process of measuring, or assessing risk and then
developing strategies to manage and mitigate the risk. In general, the strategies
employed include transferring the risk to another party, avoiding the risk, reducing
the negative affect of the risk and accepting some or all of the consequences of a par-
ticular risk (www.dictionary.com) (Figure 4.26).

Figure 4.26 Risk management


Source: McMullen, 2007
390 Green and sustainable mining

Risk assessment is the first step in the risk management process. Risk assessment
is measuring two quantities of the risk, the magnitude of the potential loss and the
probability that the loss will occur.
Much of the following is modified from McMullen, 2007.

4.6.3.1 Managing risk


The following steps should be taken to eliminate, control or reduce risk:

• An assessment must be conducted by reviewing previous incident, accident,


injury, first aid and loss records.
• These assessments must be categorized and the most serious and the most com-
monly recurring should be addressed first.
• Action must be taken to manage the identified risk by either eliminating it or
mitigating it.
• Follow-up observations and audits must be conducted to identify new risks and
ensure that no reoccurrences happen.
• Report findings effectively, and if necessary, take further corrective action.

The human element is always present in risk, and employees often knowingly put
themselves at additional and unnecessary risk for various mainly behavioural rea-
sons, including:

• To save time
• Doing the task always incorrectly but nothing adverse has happened to date
• Performing the task safely may appear inconvenient
• Incorrectly trained to do it that way
• Untrained and unaware of the hazard

The following gives some common factors in accidents:

• Risk – There is always an element of risk in any activity.


• Behaviour – Not realising the hazard that was present or believing that doing it
incorrectly was not a problem.
• Time – The majority of people involved in accidents say they were trying to
save time.
• Lack of Management – Management had not taken the time to identify hazards
and control them (with training, for example).

4.6.3.2 Risky behaviour


Risky behaviour is influenced or caused by:

• An action performed by a person that can cause harm to them or another person
• A source of possible danger that could and should be rather avoided altogether or
done differently
• Willingly exposing a person or another person to an avoidable hazard
Green and sustainable mining 391

• There is a potential loss that may occur


• Taking a dangerous chance
• Hoping for a favourable outcome rather than ensuring a favourable one based on
taking action
• Where prior knowledge exists and it is ignored

Risky behaviour can be positively influenced, as it is habit related, and habits can be
modified or changed. Leading by example and by showing that safety is always first,
poor or dangerous behaviour can be reduced, and a more positive influence can affect
the culture of a workforce and create a safer work environment. Poor or risky behav-
iour must always be corrected when observed.
Time is money in industries, and when tasks are assigned, a realistic time must
be allocated to ensure safety. Such major activities, when this is even more critical,
include:

• Most engineering projects


• Mine access development
• Equipment maintenance
• Complex task analysis
• Start-ups and shutdowns
• Major construction projects
• Rebuilds
• Coal longwall panel moves

4.6.3.3 Managing risk


Risk must be managed or it will manage the mine and create a poor work environ-
ment. The key is to be collecting, monitoring and analyzing and making decisions
based on relevant data preferably in real time.
Important related activities are:

• Monitoring (preferably in real time)


• Continuous training and upskilling
• Incident and accident analysis and recurrence prevention or mitigation
• Auditing
• Compliance (with regulatory and other guidelines and standards)
• Stakeholder relations
• Sound and transparent management

Areas that require monitoring to reduce risk include:

• Asset operation, location and movement (mainly of personnel and equipment)


• Air quality and flow
• Temperature measurement (ambient and on equipment and moving components)
• Tailings dam
• Water quality and discharge
• Any mining-induced surface subsidence
392 Green and sustainable mining

With the Fourth Industrial Revolution and the growing use of real-time monitoring,
automation, artificial intelligence and robotics, cyber security has become a new and
very important risk that needs to be effectively monitored and protected. This is cur-
rently not adequately addressed.
Mines often introduce incentive and bonus programs to encourage safe and pro-
ductive behaviour, but these are not simple and often have unplanned negative conse-
quences. These could include the following:

• Employees may try to hide work injuries.


• If the safety bonus is team dependent, an injured member may be “targeted and
harassed” if the other members lose money because of the injury.
• Injured employees may come back to work or be coerced to return prematurely so
as to reduce the apparent severity of an injury.

4.6.4 Future mining technologies

4.6.4.1 Introduction
Mining and mining technology will change in the future, and this will change and
hopefully reduce the risk.
It is certain that the mine of the future will be:

• Safer
• More environmentally friendly and sustainable
• Make full use of big data (for real-time monitoring, etc.)
• Require a more highly skilled workforce
• Make extensive use of artificial intelligence and robotics

According to an article by Husseini (2018) in Mining Technology in September, the


following new or better utilized technologies could shape the Future Mine, based on
eight industry leaders input and opinions:

• Optimizing technology using the Internet of Things (Source: Joe Carr, Director of
Mining Innovation, Inmarsat) – The clever implementation of digital technolo-
gies like the Industrial Internet of Things (IIoT) and automation could trans-
form mining, making it safer, more productive, efficient, sustainable, and
profitable, and therefore better able to take on the challenges it faces. When
we consider that over the last 15 years, the average cost of producing copper
has risen by more than 300%, while the grade has dropped by 30%, these new
efficiencies offer a cost-effective way to increase profitability. One of the
biggest areas of promise is in IIoT’s ability to transform expensive and inef-
ficient manual and mechanical processes into digital ones. IIoT technology
enables mining organisations to collect vast quantities of data about their
operations remotely and in real time through internet-connected sensors.
This data can then be acted upon and used to improve efficiency on site,
ensure a safe environment for miners and monitor the operational status of
machinery.
Green and sustainable mining 393

• Smart Mines and Artificial Intelligence (Source: Chris Mason, Director of Sales
for EMEA, Rajant Corporation) – The mining industry is a combination of brute
force and some of the most advanced scientific and mathematical processes used
in any industry. The application of technology will continue to remove people
from the brute force aspect of the business, whilst advancing the ability to find,
extract and process mined materials, quicker, cheaper and at a better rate per
tonne. Given the onerous nature of the work, the future will see mine employ-
ees focused on the business aspects of mining, such as managing a company’s
strategic relationships, and not in the field. Machines will not only be able to
operate autonomously to a pre-determined plan, but will process data them-
selves and make decisions when circumstances change and sensors detect differ-
ent conditions. Not only will the mines themselves be intelligent and all assets
connected, but the value chain from mine all the way to the ultimate user of
the materials will be connected, so that production can be planned and flexed
to meet demand and adapt to resulting changes in commodity price. Artificial
intelligence will make decisions on production and routes to market, informed
by learning from connected global trends and the real-time capabilities of the
companies’ mining properties.
• The use of Blockchain (Source: Rebecca Campbell, mining partner, White & Case
LLP) – Blockchain is being trialed by mining firms to track and trace key com-
modities, including conflict minerals and diamonds, as they move from the
mine site through the supply chain. Blockchain is a continuously growing list
of records logged in a decentralised, immutable, cryptographically secured
online ledger. All information recorded on the platform is peer-to-peer val-
idated and therefore can’t be corrupted without others quickly becoming
aware.
The recent modest recovery in mining productivity has been threatened again
as demand improves and prices recover, and as a result, the industry is under
pressure to focus on methods to improve efficiency. Naturally this falls on the
supply chain, and we believe, blockchain and smart contracts will be a key
building block to achieve this. Blockchain’s role in sustainable and transparent
supply chains could be a game changer, thanks to its ability to promote track-
ability, transparency and security through open peer to peer and incorruptible
data sharing. This will offer a new tool to monitor and confirm compliance with
sustainability and environmental ethical standards. Further to this, smart con-
tracts, with their self-executing automatism and case of replication can create
significant efficiencies for global procurement teams and help ensure regulatory
compliance. Ultimately, with more pressure on margins and inflationary costs,
as well as corporate social responsibility, this technology will be key for the
future development of the mining industry’s supply chain. It’s not a question of
if, but when.
• The utilization of new metals (Source: William Adams, head of research, Metal
Bulletin) – While base and precious metals mining are mature markets, we are
witnessing the start of a new era for lithium and cobalt production as demand
for these two materials undergoes exponential growth from lithium-ion batteries.
The compound annual growth rate for Electric Vehicles (EVs) uptake over the
next decade is expected to be around 27%, but already growth rates are much
394 Green and sustainable mining

higher than this, albeit off a low base. China’s EV sales are rising by around 80%
year-on-year and in Q2 in Europe, battery EV registrations were up 45.5% com-
pared with Q2 2017. Imagine what sales growth rates will be when battery-only
EVs are cheaper than internal combustion engine vehicles. Following the demand
shock in 2015, when China put its weight behind EVs, that saw a rapid run up in
lithium and cobalt prices, we are now seeing the supply response that has led to
price falls this year, but with the end use market for lithium-ion batteries already
growing rapidly, but still at the early stages of the innovation ‘S’-curve, lithium
and cobalt producers are going to work hard to bring enough supply on in a timely
manner. Battery cell requirements for EVs, passenger and commercial, are fore-
cast to grow from around 70GWh in 2017, to 1,600GWh in 2030. For any supply
chain to keep up with such a rate of growth for an extended period of time will be
exhausting, but it will require the lithium and cobalt markets to grow from small
markets to significant markets in double quick time.
• Drone Technology (Source: Joe Carr, director of mining innovation, Inmarsat) –
The mine remains a uniquely hazardous and inconvenient workplace. Mines
often have to pay higher wages for remote workers, in addition to high transport
and accommodation costs. Another challenge is the impact of high workforce
turnover caused by the ‘fly-in fly-out’ lifestyle. IIoT and automation offers a way
around all the risks and expense inherent in employing people in these locations,
while bringing the precision and bandwidth of technology to the heart of remote
mining operations, all whilst improving productivity.
For example, Freeport-McMoRan is already using drones to create steeper pit
slope angles in its mines, reducing the stripping ratio and amount of waste rock
hauled before ore can be extracted. These drones not only scan the mines from
perspectives that are dangerous and near inaccessible to humans, they also instan-
taneously communicate any information they pick up. This makes for a more
rapid and detailed analysis of the mine slopes without having to deploy highly
skilled geologists or geotechnical engineers into an inherently hazardous environ-
ment or affecting production by closing haul roads. With machines becoming pro-
gressively more capable of acting with little manual intervention, a future where
adaptable and autonomous machines carry out the on-site, operational tasks of
mining while human employees monitor them remotely looks probable and highly
profitable.
• Overcoming the skills gap (Source: Ed Harrington, director, The Open Group’s
Open Process Automation Forum) – Critical infrastructure industries such as
mining are struggling to attract and retain the right technological capabilities.
Ultimately, the main barrier for graduates entering these industries is propri-
etary outdated technologies, which demands time-consuming and expensive
training and limits future job prospects. As a result, the job market stagnates
and older generations are the only people with the knowledge of how specific
systems work.
However, the operational technology and automation space is moving to the
same compute stack–and if we can streamline software languages, these various
parties will soon begin to speak the same language. This will reduce costs and
streamline effectively so that there are enough people within an organisation
and the wider industry sharing the same knowledge and skillsets. The alternative
future is one where we see an ever-increasing skills gap.
Green and sustainable mining 395

Looking ahead, companies will need to agree on technical standards that are
open, based on common languages. This means process control systems in one
organisation are compatible with those from another. This will not only make it
easier and cheaper for existing staff to replace and repair control systems, it will
also be easier for new talent to be attracted to and retained within the mining
sector.
• The future of mining: preserving authenticity in the gold supply chain (Source: Kai
C Chng, CEO and founder, Digix) – Traditionally, gold investments have taken
the form of share investments into gold exchange-traded funds or acquisi-
tion of physical gold bars, which then need to be stored. The introduction of
blockchain technology and its inherent attributes brings about many advan-
tages to the transfer of gold. The public ledger and full public access to all
traceability and authentication documentation can minimise fraud risks that
are present in what would otherwise be just a physical supply chain of asset
custody.
• Increased technology underground (Source: James Woodall, co-founder and CTO,
Intoware) – Now that intrinsically safe devices are more prevalent, we’ll see more
technology actually down in the mines and the savvy businesses will be those that
start to take advantage of this early. This will include improving on processes
such as digital audit checks, where safety and quality control remains paramount.
But we’re likely to see an increase in areas like in-location video conferencing and
digital product tracking.

4.6.4.2 A major mining company’s vision


Rio Tinto is one of the leaders in the industry, investing heavily in research associated
with the “Future Mine”. Their approach is (Be Smart website-2021 - www.riotinto.
com/en/about/innovation/s):

Data is one of our most valuable assets.


“Every day our automated drills, trucks, shovels, conveyors, trains and ships
produce huge amounts of valuable data. By combining this data with clever ana-
lytics, artificial intelligence, machine learning and automation, we are making
our business safer and more productive.
We have also set up Centres of Excellence focused on analytics, automation,
asset management, energy and climate change, orebody knowledge, underground
mining, surface mining and processing. They bring together our foremost tech-
nical experts to work with our operations around the world. These smart mining
teams help us make the right technical judgements and decisions to help us man-
age our major hazard risks and assure the safety of our assets and operational
excellence.

Specific technologies they are testing or implementing include:


• The use of virtual reality (VR) for training – Figure 4.27 is an example of how Rio
Tinto is using VR training in one of their smelters.
• Mine Automation System (MAS) – According to Rio Tinto: “MAS essentially oper-
ates like a network server application, pulling together data at 98% of our sites,
396 Green and sustainable mining

Figure 4.27 A
 VR plan of their aluminium smelter is used for training and to help avoid or handle
safety incidents (Rio Tinto)

and mining it for information. MAS provides this information in a common for-
mat, using sophisticated algorithms. It can be displayed visually using RTVis™
(Rio Tinto Visualisation) or through more conventional operational type dash-
boards with graphs, charts and tables.” Figure 4.28 shows how the system can be
used to show the location of assets underground.

Figure 4.28 An example using MAS (Rio Tinto)


Green and sustainable mining 397

• Tracking equipment and data mining – According to Rio Tinto: “Satellites stream
oceans of data telling us everything from the position of a truck to where our next
discovery might be. We have built computer systems that analyze data and make
decisions in microseconds”.
• RTVis: A 3D visualization tool – According to Rio Tinto: “We built RTVis with a
3D gaming engine to help us see inside our operations. We can fly over a site and
dive down to the detailed information we need. We can hover over an excavator,
follow a haul truck or examine an orebody.”The software brings together geology,
geotechnical, drill and blast, production and planning, and visualizes surface and
sub-surface features. There are also various analytics tools to help us make sense
of our data and information for better decision-making. An example of how this
technology can be used is shown on Figure 4.29.

4.6.4.3 Space mining


The following is taken from “Mines of the future major findings report”, published by
the Society of Mining Professors (SOMP) (2019).
As space lacks water, the first off-earth mining operation will most probably be the
extraction of water from the moon surface to be used as a rocket fuel to sustain life in
space.
In the last two years, the focus has been on asteroids and then Mars, but focus
recently moved to the moon. However, NASA considers that the moon is just a step-
ping-stone, not an alternative to Mars. The moon is rich in resources, and it is esti-
mated there are more than 6 billion tonnes of water on the moon.
The non-existence of a market in space is stopping potential research activities.
However, recently, United Launch Alliances (ULA) announced that they will be

Figure 4.29 A 3D surface mine visualization model (Rio Tinto)


398 Green and sustainable mining

purchasing fuel (H, O) for US$3,500/kg in Low Earth Orbit and for US$500/kg on the
moon. This would certainly increase the interests of potential business investments
and research opportunities.
Operations will have to be conducted by a robotic equipment fleet. Numerous off-
earth activities have the potential to expand and stimulate the global economy, such
as construction materials for off-earth structures, solar power stations to reduce
greenhouse gas emissions while expanding energy access, space tourism, propel-
lant depots and resources for human habitation. Various risks to these commercial
ventures include legal risks, business cases and hazards to spacecraft such as space
debris.
Aside from the material from SOMP above, the following is relevant. The concept
of “space mining” began to develop globally in the early ’90s but caught momentum
in the US on November 25th, 2015, when President Obama signed the so-called “Space
Law”, which was approved by the US Congress.
A conceptualized space mining camp is shown on Figure 4.30.
The UN in 1967 passed the “International Treaty on Outer Space” that bans indi-
vidual countries from claiming exclusive rights to anything in outer space.
In 1979, the UN passed the Moon Treaty, which specifies that all activities
on the moon (and other celestial bodies) are carried out in the interest of peace
and with due regard to the interests of all parties and other states. The moon is
also to remain demilitarized with no nuclear weapons or other weapons of mass
destruction.
It is hoped that both agreements are in fact honoured by all nations in the future.

Figure 4.30 Sketch of a possible space mining camp


Source: NASA
Green and sustainable mining 399

4.7 Mining valuation and financing


Authors: Lilford, E.V. & Spearing, A.J.S.
The reliable estimation and public reporting of resources and reserves is absolutely
critical for the mine operator, the shareholders, the regulators, the public and any
other stakeholder. Clearly, geologists play a key important role in this.
Mineral exploration consists of the following main activities:

• Determining the mineral of interest


• Area selection
• Target definition
• Ground disturbance activities
• Resource evaluation
• Reserve definition
• Environmental assessments (EA)
• Stakeholder engagement
• Environmental Social & Governance (ESG) considerations
• Preliminary and detailed economic assessment
• Trial/full extraction

4.7.1 Mineral resources


According to the Joint Ore Reserves Committee (JORC, 2012) of Australasia:

A ‘Mineral Resource’ is a concentration or occurrence of solid material of eco-


nomic interest in or on the Earth’s crust in such form, grade (or quality), and
quantity that there are reasonable prospects for eventual economic extraction.
The location, quantity, grade (or quality), continuity and other geological charac-
teristics of a Mineral Resource are known, estimated or interpreted from specific
geological evidence and knowledge, including sampling. Mineral Resources are
sub-divided, in order of increasing geological confidence, into Inferred, Indicated
and Measured categories defined as follows:

• ‘Inferred Mineral Resource’ is that part of a Mineral Resource for which quan-
tity and grade (or quality) are estimated on the basis of limited geological evi-
dence and sampling. Geological evidence is sufficient to imply but not verify
geological and grade (or quality) continuity. It is based on exploration, sam-
pling and testing information gathered through appropriate techniques from
locations such as outcrops, trenches, pits, workings and drill holes. An Inferred
Mineral Resource has a lower level of confidence than that applying to an
Indicated Mineral Resource and must not be converted to an Ore Reserve. It is
reasonably expected that the majority of Inferred Mineral Resources could be
upgraded to Indicated Mineral Resources with continued exploration.
• ‘Indicated Mineral Resource’ is that part of a Mineral Resource for which
quantity, grade (or quality), densities, shape and physical characteristics are
estimated with sufficient confidence to allow the application of Modifying
Factors in sufficient detail to support mine planning and evaluation of
the economic viability of the deposit. Geological evidence is derived from
400 Green and sustainable mining

adequately detailed and reliable exploration, sampling and testing gathered


through appropriate techniques from locations such as outcrops, trenches,
pits, workings and drill holes, and is sufficient to assume geological and grade
(or quality) continuity between points of observation where data and samples
are gathered. An Indicated Mineral Resource has a lower level of confidence
than that applying to a Measured Mineral Resource and may only be con-
verted to a Probable Ore Reserve.
• ‘Measured Mineral Resource’ is that part of a Mineral Resource for which quan-
tity, grade (or quality), densities, shape, and physical characteristics are estimated
with confidence sufficient to allow the application of Modifying Factors to sup-
port detailed mine planning and final evaluation of the economic viability of the
deposit. Geological evidence is derived from detailed and reliable exploration,
sampling and testing gathered through appropriate techniques from locations
such as outcrops, trenches, pits, workings and drill holes, and is sufficient to
confirm geological and grade (or quality) continuity between points of observa-
tion where data and samples are gathered. A Measured Mineral Resource has
a higher level of confidence than that applying to either an Indicated Mineral
Resource or an Inferred Mineral Resource. It may be converted to a Proved Ore
Reserve or under certain circumstances to a Probable Ore Reserve.

Resource evaluation is:

• Used to quantify the grade and tonnage of the orebody


• Generally achieved after completing drilling on a pre-determined, usually regular
grid pattern
• Potentially augmented through additional holes in “areas of interest” being com-
pleted (e.g., major fault zones, etc.)
• Often extended to incorporate holes for geophysical probing, depending on the
orebody (e.g., sulphides)

4.7.2 Mineral reserves


According to JORC (2012):

An ‘Ore Reserve’ is the economically mineable part of a Measured and/or Indicated


Mineral Resource. It includes diluting materials and allowances for losses, which
may occur when the material is mined or extracted and is defined by studies at Pre-
Feasibility or Feasibility level as appropriate that include the application of Modifying
Factors. Such studies demonstrate that, at the time of reporting, extraction could rea-
sonably be justified. The reference point at which Reserves are defined, usually being
the point where the ore is delivered to the processing plant, must be stated. It is impor-
tant that, in all situations where the reference point is different, such as for a saleable
product, a clarifying statement is included to ensure that the reader is fully informed
as to what is being reported. Ore Reserves are further divided into:

• A ‘Probable Ore Reserve’, the economically mineable part of an Indicated,


and in some circumstances, a Measured Mineral Resource. The confidence in
the Modifying Factors applying to a Probable Ore Reserve is lower than that
applying to a Proved Ore Reserve.
Green and sustainable mining 401

• A ‘Proved Ore Reserve’, the economically mineable part of a Measured


Mineral Resource. A Proved Ore Reserve implies a high degree of confidence
in the Modifying Factors.

Reserves are:

• A converted (more reliable) mineral resource after more detailed exploration


• More intensive and technical than resource evaluation and aims at statistically
grading the orebody and geometry
• Inclusive of metallurgical testing and rock mechanics mine design and support
assessments
• Aimed to produce a full bankable feasibility study proving the economic viability

4.7.2.1 Resource and reserve reporting


The requirements for determining and reporting mineral resources and reserves are
defined in countries, examples of which are:

• The JORC from Australia.


• Committee for Mineral Reserves International Reporting Standards (CRIRCSO),
initially started by the Council of Mining and Metallurgical Institutions.
• The SME Guide for Reporting Exploration Results, Mineral Resources, and
Mineral Reserves in the US.
• The South African Mineral Resources Code (SAMREC).
• The NAEN Code is a Russian Code for the Public Reporting of Exploration
Results, Mineral Resources and Reserves.
• United Nations Framework Classification for Fossil Energy and Mineral
Resources (UNFC) was developed in 1997 as an initiative of the United Nations
Economic Commission for Europe.
• The Chinese State Mineral Resource and Reserve System is based on the UNFC
classification.

One of the more popular systems is the JORC Code (first established in 1971 and the
first code published in 1989); it is based on minimum principles and standards mainly
for consistent and accurate reporting rather than a prescriptive code, is well accepted
globally and is frequently reviewed and updated based on a wide consensus build-
ing method by all stakeholders. The report based on the JORC Code must always be
signed off by a responsible person.
Based on the JORC Code (2012):

A ‘Competent Person’ is a minerals industry professional who is a Member or


Fellow of The Australasian Institute of Mining and Metallurgy, or of the Australian
Institute of Geoscientists, or of a ‘Recognised Professional Organisation’ (RPO),
as included in a list available on the JORC and ASX websites. These organisations
have enforceable disciplinary processes including the powers to suspend or expel
a member. A Competent Person must have a minimum of five years relevant expe-
rience in the style of mineralisation or type of deposit under consideration and in
the activity which that person is undertaking. If the Competent Person is preparing
402 Green and sustainable mining

Figure 4.31 The value and reliability of the orebody during exploration
Source: H.Z. Harraz, 2014, JORC Presentation, August

documentation on Exploration Results, the relevant experience must be in


exploration. If the Competent Person is estimating, or supervising the estima-
tion of Mineral Resources, the relevant experience must be in the estimation,
assessment and evaluation of Mineral Resources. If the Competent Person is
estimating, or supervising the estimation of Ore Reserves, the relevant experi-
ence must be in the estimation, assessment, evaluation and economic extraction
of Ore Reserves.

Figure 4.31 compares the value and reliability of the geological model during exploration.
Figure 4.32 (after JORC, 2012) shows the relationship between mineral resources
and ore reserves based on increasing geological data.

4.7.3 Mineral economics


Mineral economics is the study and application of the economic and policy issues
impacting the production, management, control and finance associated with the dis-
covery, development, exploitation, marketing and rehabilitation of operations, assets
and companies within the minerals sector.
Although not exclusively, and notably for producing assets, mineral economics typ-
ically depends on the supply and demand of mineral commodities. It is a complex
topic that involves:

• Mine grade, production and total costs, including capital development costs
• The finite life (non-renewable) of a mine asset (unlike, for example, agriculture,
which should be renewable, climate permitting)
Green and sustainable mining 403

Figure 4.32 The relationship between exploration results (JORC, 2012)

• Mineral and related global prices, including appropriate currency exchange rates,
that are related to supply and demand, incorporating substitution, and influenced
by any significant recycling and relevant technology changes
• Financial control measures
• Relevant regulations and policies
• Socio-political issues, including sovereign risks
• ESG factors, including mainly local communities as well as climate change
impacts
• Global markets, notably in securities exchanges and Over The Counter (OTC)
derivative markets

For non-producing assets, mineral economics offers a more complex arena in which
properties and assets hosting or associated with commodities are evaluated outside of
using conventional evaluation tools.
According to Jain (2020), and excluding specific markets and market influences,
mineral economics is a unique branch of applied economics because:

• An orebody is hard to find and then economically justify, and finding an ore-
body is becoming harder as the higher grade and more accessible ore deposits are
becoming exhausted.
• An orebody is finite and non-renewable.
• Mineral deposits are full of uncertainties especially around orebody size and
grade.
• It takes expending considerable time and money between exploration and the
start of exploitation.
• An orebody location is geology-defined, not country-defined.

4.7.3.1 Mineral property valuations


Mining assets are mainly made up of mine prospects (exploration permits) and oper-
ating mines (mining or exploitation permits).
404 Green and sustainable mining

Mine valuation basically involves the weighing up of all the financial aspects of a mine
in order to place a reliable present value on the ore reserves and resources. The value of
a mineral deposit must be reliably developed and updated as markets change and more
technical and techno-economic data become available during the mine’s operating life.
In general, there are four main purposes supporting mineral project valuations
(Frimpong, 1992), to:

• Indicate technical, economic and operational guidelines to achieve the optimal


exploration and exploitation programmes of a project.
• Ascertain the viability and the associated value and the inherent uncertainty of a
mineral project.
• Facilitate decision-making relating to the development, acquisition, project financ-
ing, regulatory requirements and taxation considerations surrounding a project.
• Provide management with adequate tools to incorporate and act on the inherent
flexibility in a project to improve operating standards and procedures and to con-
trol operating variances.

Beyond the above four main purposes, additional reasons include valuations being
required (Lilford, 2012):

• To accommodate transactions culminating in mergers and acquisitions, as well as


mineral asset disposals, conducted explicitly or relatively.
• To comply with legislated requirements such as local party (community or com-
pany) or government participation in that specific country’s minerals sector.
• To facilitate litigation proceedings.
• To reflect stock exchange regulated fairness and reasonableness opinions or inde-
pendent expert reports (for minority shareholder protection).
• For expropriation considerations resulting from changing legislation or dispos-
session claims (including mediation and arbitration proceedings).
• To evaluate financial security for debt provisions, indicating the robustness of a
project and accommodating specific lending ratios to provide adequate comfort
(financial covenants) to a lender.
• For insurance claims.
• For accounting purposes.
• For Initial Public Offerings (IPOs) or new listings and other equity-raising exer-
cises (rights offers, private placements, management buyouts, etc.).

To determine the explicit value of a mineral project can be challenging, and developments
in the global resources industry have attempted to regulate and somewhat standardize
the methodologies employed to valuate mining and exploration opportunities. While the
in-situ minerals cannot be traded (bought and/or sold), they can be once they are mined
and processed or refined. However, companies can still trade exploration or mining per-
mits associated with those mineral properties containing in-situ deposits. Arising from
this, accurately determining a fair value of an exploration or mining permit is important,
and according to CIMVal (2019), is dictated by three important factors:

• The independence to transact (between two companies or two parties)


• The valuation and assessments being conducted at arm’s length
• The parties being under no obligation to transact
Green and sustainable mining 405

Table 4.1 Valuation approaches and application

Valuation approach Application

Cost Approach Appropriate for early-stage mineral property exploration assessments,


where some exploration expenditure has been incurred on the property.
Applies to exploration permits and assets under exploration.
Market Approach Appropriate for early-stage mineral property exploration assessments,
where little to no expenditure has been incurred on the property, but
some general information is either known or can be inferred. Applies to
exploration permits and assets under exploration.
Income Approach Appropriate for mineral properties that have undergone some exploration
activities such that at least a Measured or Indicated Resource is known,
but ideally a Reserve has been identified, and a cash flow analysis can be
compiled. This applies to all mining license applications or holders.

It is important to recognize the nomenclature used that applies to mineral property


valuations. Specifically, a valuation approach defines the broad category under which
valuation methodologies are grouped.
Mineral property valuations fall into one of three Valuation Approaches, the Cost,
Market and Income Approaches (Stephenson, 2017), as shown in Tables 4.1 and 4.2
(modified from Lilford, 2012).
The appropriate valuation approach and corresponding methodology to be used
depends on the degree of information available about the property, as highlighted in
Table 4.2. If the property has defined Reserves on it, that property should be valued
using the Income Approach and probably a Discounted Cash Flow (DCF) method,
including a Net Present Value (NPV) determination. This is because a defined Reserve,
by definition (JORC, 2012), has had an economic assessment conducted on it in order to
be categorized Reserve, implying that some form of extraction plan with costs and cap-
ital expenditure has been determined, making it possible to identify an income stream.
By contrast, the Cost Approach is best suited to properties over which only his-
torical exploration expenditure information is known, or for which terms have been
drawn up for a JV or an earn-in or farm-in arrangement, but for which there is no
identifiable income yet.
The Market Approach is best suited to properties that have had no formal invasive
(ground disturbing) exploration work conducted on them, but which properties are
likely to be mineralized based on geological or proximity factors.

4.7.3.2 Cost approach and associated methodologies


The Cost Approach and the valuation methodologies grouped within this approach
may be applied to properties that have undergone or may still be undergoing explora-
tion, but for which no cash flows are being generated.
The most widely used valuation methodology under the Cost Approach is the
Multiples of Exploration Expenditure method.
MULTIPLES OF EXPLORATION EXPENDITURE (MEE)

The value of an exploration project can be determined based on a multiple of the


actual exploration expenditure (MEE) that has been spent or is committed to be
spent (budgeted) over the forecast exploration tenure. It is an appropriate method
406 Green and sustainable mining

Table 4.2 Valuation approaches, methodologies and data description

Approach Method Summary of method and required data

Cost Multiple of Exploration Appropriate past exploration costs and budgeted/projected


Expenditure (MEE) future costs, inflated/deflated to current money terms,
multiplied by the appropriate multiplier.
Data required: past exploration expenditure and associated
expenditure dates, or budgeted exploration expenditure
forecasts
Farm-in/Farm-out Amounts spent over time to earn equity interests correlates
analysis to overall value, in real money terms.
Data required: the terms of the buy-in (amounts and timing)
or sales, as well as the equity amounts attributed to each
farm-in or farm-out date
Market Dollar per hectare A rudimentary method that equates a project area’s value
to its size against previous transactions in the area.
Data required: size (area) of mineral property, and relevant
industry transactions to obtain a comparative $/ha rate
Market capitalization A listed company’s market capitalization divided by its ounces
per unit or tonnes ($/oz or $/t of contained mineral) can be used as a
proxy for other project values.
Data required: a relevant listed company from which a $/oz,
$/t or other metric can be determined
Comparable Similar properties reflect similar values, matrix includes depth
transactions using below surface to top of mineralization, grade in situ, geological
an appropriate certainty and proximity to useful infrastructure.
matrix Data required: depth of top of orebody below surface, in-situ
mineral grade, geological descriptor (JORC categorization),
proximity to useable infrastructure and a relevant comparative
value-table (matrix of transactions)
Option agreement/ Required participation expenditure in proving up or developing a
JV terms project by the incoming party, acquiring additional equity as
expenditure occurs.
Data required: the terms of the option agreement or JV
identifying the costs to participate at that stated percentage
equity holding
Income DCF NPV performed on calculated free cash flows (FCF) over a
project’s life.
Data required: details around tonnage forecasts, grades,
recoveries, capital expenditure, operating costs, etc., so that a
detailed cash flow can be drawn up. The applicable discount rate
should be provided too so that a NPV can be determined
Real Options Based on the NPV method, various management choices can be
analyzed in the valuation to determine a more accurate value.
Data required: the terms of the option
Decision Tree Similar to Real Options, where value outcomes are dependent
(binomial and upon decisions made at the decision points.
multinomial) Data required: similar data to options, but may include
probabilities of outcomes

provided that only previous and planned expenditure relevant to the explora-
tion is included, and that the quality of the exploration work is appropriate and
acceptable.
The attributable exploration expenses shall only include expenses directly related to
exploration and must exclude other, unrelated expenses, as shown in Table 4.3.
Green and sustainable mining 407

Table 4.3 Included and excluded costs and fees for the MEE method

Directly Related Expenses (included in MEE) Indirectly Related Expenses (excluded from MEE)

Direct and indirect drilling expenses Service fees of electronic users


Trenching and sampling costs The sales price of the selection process
Appropriate and applicable salaries (e.g., drilling Application service fee, directors’ fees (unless a
crew, field geologist managing the drilling director is on-site managing the exploration)
programme, personnel employed to assist
drilling, etc.)
Laboratory and assay costs Annual license fees
Service fees for plan approvals Any other costs or fees that are not directly
attributable to the exploration activity
Other services directly related to the
exploration outcome

The historical costs that have been incurred on exploration activities must be inflated to
current day monetary terms at the appropriate annual rate (usually the Consumer Price
Index-CPI), and any future planned and budgeted exploration costs must be discounted
(deflated) to current day money terms at the forecast annual inflation (CPI) rate.
The actual multiple (multiplier) for the MEE must be selected as follows. If mineral-
ization has been identified on the property, the multiple to be applied to the historical
inflated and future (budgeted) deflated exploration expenditure must be determined. If
the exploration area is determined to be void of mineralization, or contains low levels of
mineralization that are likely to render the area uneconomic for extraction purposes, the
value of the permit will be zero. However, if mineralization has been identified, the mul-
tiples to apply, dependent on the geological certainty of the deposit, follows in Table 4.4.

Table 4.4 Applicable multiples dependent on geological certainty

Measured &
Proven Reserve Probable Reserve Indicated Resource Inferred Resource Blue-sky

3 2.4 1.8 1 0.5

In most cases of an exploration site, it is highly unlikely that a Proven or Probable


Reserve exists. Most likely, the exploration target will fall into the Inferred Resource
category unless information proves otherwise.
Example: Exploration Permit EP1 was granted five years ago, with exploration com-
mencing the following year. Exploration has been conducted over the last four years
on the property. The submitted exploration report clearly shows the exploration
expenditure amount that was incurred each year. Exploration activities proved up the
deposit to an Indicated Resource geological category (Table 4.5).

4.7.3.3 Market approach and associated methodologies


As with the Cost Approach, the Market Approach and its associated valuation
methodologies are generally not appropriate to use for mining (exploitation) opera-
tions but rather can be used to value exploration properties. This approach may be
used in preference to the Cost Approach, but only where the available information
408 Green and sustainable mining

Table 4.5 Multiples of exploration expenditure valuation result - example

Exploration expenses included


in exploration report Escalated Total Value
Year (US$’000) CPI (US$’000) Multiple (US$’000)

–4 625 2.60% 699


–3 745 2.20% 812
–2 800 1.70% 853
–1 710 2.30% 744
0 920 2.50% 943
Amount Year (0) 4,051 1.8 7,293

surrounding the project lends itself preferentially to this approach. This will occur
in the event when no actual exploration expenditure has been incurred on the pro-
ject area yet so that applying the MEE method cannot occur, or possibly that the
exploration expenditure was incurred a significantly long period of time ago and is
now completely inappropriate.
The Market Approach is based on comparing mineral project transactions that
have taken place in an appropriate or proximate market, and by applying these com-
parative unit ($/ha, $/oz, $/t, etc.) values to the mineral property being valued.
The mineral properties must be similar in material respects, including in the type of
commodity (gold, copper, lithium, zinc, etc.), orebody layout (depth, thickness, shape,
etc.) and approximate size.
There are at least three valuation methods within the Market Approach: the dollar
per hectare method, the Techno Economic Matrix (TEM) method and the market
capitalization per unit method. Each will be considered further.
DOLLAR PER HECTARE

This method simply suggests that previous market transactions for exploration prop-
erties will not differ much from new transactions when measured on a $/ha (dollar per
hectare) basis. This method applies to cases in which very little information is known
about the exploration target.
TECHNO ECONOMIC MATRIX (TEM)

The Lilford TEM valuation method (Lilford, 2004, 2012) is applied when specific key
parameters are known such that they can be awarded points according to the pre-de-
termined matrix. The four key assessment parameters for the TEM method, shown in
Table 4.6 with corresponding points attributions, are:

• The vertical depth to the top-of-mineralization below surface


• The geological certainty categorization (Proven, Probable, Measured Indicated,
Inferred, Blue Sky). It is improbable that Proven or Probable Reserves will be
known at this stage.
• The in-situ grade (e.g., g/t for gold, % for copper, etc.)
• The proximity of that mineral project to existing mining and/or other essential
infrastructure (measured in kilometres)

The “$/ha applicable” column in Table 4.7 which is presented solely for indicative
purposes, must be determined (updated) based on transactions that have occurred in
Green and sustainable mining 409

Table 4.6 Valuation matrix for gold properties and gold equivalence

Depth below surface Resource category In-situ grade Proximity

km points points g/t points points

0.00–0.30 0 Proven 0 0–1 7 Contiguous (<5 km) 1


0.30–1.00 1 Probable 1 1–2 6 Proximal (5–30 km) 2
1.00–2.00 2 Measured 2 2–3 5 Semi-proximal (30–80 km) 3
2.00–4.00 3 Indicated 2 3–4 4 Remote (80–150 km) 4
>4.00 4 Inferred 3 4–5 3 Distal (>150 km) 5
Blue sky 4 5–6 2
6–8 1
8+ 0

the country or in surrounding countries for the specific commodity to be valued, over
a fairly recent time period. Importantly, the value column ($/ha applicable column)
must be continually updated based on any recent transactions, or at least it must rec-
ognize and factor in time adjustments (as measured by CPI or some other inflation
factor).
The “$/ha applicable” column can be developed for commodities other than gold,
based on actual transactions of properties or licenses containing that mineral. If a
separate matrix cannot be drawn up due to insufficient information for a specific
non-gold property, possibly due to a small number of transactions occurring in that
commodity, a gold-equivalence, or other-commodity-equivalence calculation can be
performed so that the available matrix can be applied.
Gold Equivalence Example: As shown in Table 4.8, exploration License No. EL2
hosts three identified mineral discoveries: tin at 330m below surface, lead at 110m
below surface and zinc at 50m below surface. The size of the mineralized area is
500 hectares, and the distance to supporting infrastructure in the area is 85 km.
Minimal exploration results infer that the mineral grades are zinc at 1.4%, lead at 3.6%

Table 4.7 Determination of applicable value rating

Points summed Attributable rating Attributable rating $/ha applicable

1 1 1 2,000
2 2 3 1,857
4 3 3 1,592
4 1,459
6 4–5 5 1,327
6 1,194
9 6–7 7 1,061
8 929
11 8–9 9 796
10 663
13 10–11 11 531
12 398
15 12–13 13 265
14 133
17 14–15 15 10
+17 16 16+ 1
410 Green and sustainable mining

Table 4.8 Gold grade equivalence calculation

Parameter Zinc (Zn) Lead (Pb) Tin (Sn)

Spot Price ($/t) 2,800 2,200 29,800


Spot Price ($/kg) 2.80 2.20 29.80
In-situ grade (%) 1.4 3.6 4.1
Converted grade (g/t) equivalent 14,000 36,000 41,000
Commodity price ratio 0.000051 0.000040 0.000545
Gold equivalent grade (g/t) 0.72 1.45 22.35

and tin at 4.1%, all in situ. To determine the gold-equivalent grades for the three min-
eralized occurrences, assuming a spot gold price of US$1,700/oz, follows in Table 4.8.
The matrix is used as follows:
• Assign the necessary categories from Table 4.6 to the mineral project.
• Attribute the points associated with the assigned parameter according to the table.
• Sum the attributed points.
• Take the attributed points summed to Table 4.7 and assign the corresponding
attributable rating and then the attributable unit value ($/ha).
• Multiply this unit value ($/ha) by the hectares containing mineralization.
Note: Table 4.7, and specifically the “$/ha applicable” column, is purely for illustrative
purposes. An accurate table will need to be drawn up from available mineral property
transactions in the appropriate jurisdiction(s).
Example applying the matrix: For the example given above to calculate the gold
equivalence, the value of the project can be determined. In the example and using the
tables above following mineral property value can be determined (Table 4.9).
MARKET CAPITALIZATION PER UNIT CONTENT OR PER UNIT ANNUAL PRODUCTION

For stock exchange-listed resource companies, the market capitalization of that com-
pany divided by the in-situ Reserve units (or Reserve and Resource units), or other
measure of commodity volume (e.g., tonnes, ounces, kilograms, etc.), gives the valuer
another tool to determine a value. A listed company’s market capitalization per unit
rating can be applied to a similar, non-listed company’s asset.
The market capitalization per annual production unit (e.g., tonnes, ounce) is a var-
iation of the market capitalization per unit method for exchange-listed and operating
mining assets. It simply relates a known value to an asset for a value that is being
estimated based on production volumes.

Table 4.9 Equivalence and matrix valuation example

Project Area: 500ha Metric Zinc (1.3%) Lead (3.5%) Tin (5.2%)

Depth b.s. metres 50 (0 points) 110 (0 pts) 330 (1 pt)


Geo certainty Inferred (3 pts) Inferred (3 pts) Inferred (3 pts)
Grade equivalent g/t Au 0.72 (7 pts) 1.45 (6 pts) 22.35 (0 pt)
Proximity 85 km (4 pts) 85 km (4 pts) 85 km (4 pts)
Total Points 14 13 8
Unit Value $/ha 398 531 1,327
Value (each) $’000 199.0 265.0 663.5
TOTAL Value $’000 1,127.5
Green and sustainable mining 411

Example of market capitalization per unit content: A listed company has declared that
it has rights to 750,000 tonnes of contained copper in situ (Reserves and Resources).
This listed company has a market capitalization of $140 million (say, 400 million
shares in issue and trading at 35 cents per share). Its value per unit of content is there-
fore $186.67 per tonne of contained copper.
Therefore, a similar property with, say, 250,000 tonnes of copper Reserves and
Resources would be attributed a value of $46.67 million (=250,000 × 186.67).
FARM-IN/FARM-OUT (GRADUAL ACQUISITION/DISPOSAL LINKED TO PAYMENT MILESTONES)

For one party subject to a farm-in agreement, the other contracting party is effectively
subject to a farm-out agreement, with the terms being equal.
This arrangement provides a mechanism for the acquirer to secure increasing per-
centage stakes in a project based on pre-determined milestone targets. The criteria are
usually milestone payments incurred for performing exploration activities (drilling,
assaying, feasibility studies, etc.). A farm-in agreement will typically allow for the
transfer of parcels of shares from the seller to the buyer based on specific expendi-
tures. The terms and conditions associated with how a contributing company can earn
into the equity or ownership of a minerals project or exploration target or associated
company will provide sufficient information to conduct a valuation of the whole pro-
ject based on the amount of money to be spent in acquiring that initial or additional
equity. The agreed expenditures corresponding to specific equity amounts must all be
discounted to real money terms based on forecast CPI rates.
JOINT VENTURE (JV)

JVs generally entail one company buying into the ownership of another company’s
assets or into the equity of the holding company itself. JVs are predicated on a joint
venture agreement (JVA) between two parties typically, with the JVA disclosing all
of the terms of the ultimate partnership. The agreement will clearly state how much
money is to be spent on the project or paid to the vendor, when that expenditure will be
incurred and what rights are attached to that expenditure. In most cases, the expendi-
ture in a JV is associated with increasing participation rights (generally equity) to the
party spending the money.
Example: JV Case 1. Two JV companies, company A and B, buying a mining right
from company C.
If two companies agree, through a JV, to acquire a mining right from an unrelated
company, or agree to acquire that unrelated company itself, the value of the transac-
tion is the sum of the values associated with each JV partner. That is, each company A
and B will cover its pro-rata share of the value of the company or asset being acquired.
By way of an example, if Company C is valued at $30 million, or if its asset is val-
ued at $30 million, and Companies A and B have agreed through a JV to acquire
Company C or its asset equitably (i.e., 50%:50%), then Company A will contribute $15
million and so will Company B. That is, Company A’s value will increase by at least
$15 million and Company B’s value will also increase by at least $15 million.
If the JVA states that Company A participates at 70% of the JV and Company B
at 30%, then Company A’s value (participation cost) will be at least $21 million and
Company B’s value (participation cost) will be at least $9 million.
Example: JV Case 2. Company A and B forming a JV and company A buying into
a mining right already owned by Company B.
412 Green and sustainable mining

If Company A and B form a JV and the terms are that Company A acquires pro-
rata rights to Company B’s asset, then it is deemed that Company A is the buyer and
B is the seller of a portion of those rights. The attributable value is dependent on what
percentage is acquired for what expenditure.
By way of an example, if Company B owns mining rights and Company A is to pay
Company B $4 million for a JV participation of 50%, then the total (100%) value of the
rights is $8 million. Alternatively, if Company A must spend $3 million in exploration
to acquire 30% in the JV, then the total value of the right will be $10 million.
A JVA may dictate expenditure milestones for specific equity participations amounts.
When expenditure occurs over different periods, all dollar amounts must be inflated or
deflated to real money terms by the prevailing CPI over that period.
Example: JV Case 3. Company A and B form company C, which buys a mining
right already owned by Company B (similar to Scenario 2 but here a new company C
is being formed).
If Company A and B create a new JV company, Company C, for the purposes of
acquiring the assets of Company B, then it is deemed that Company B is a seller and both
Company A and Company B are buyers (through Company C). This may be inefficient
for Company B, but the reasoning may be sound. Under this scenario, Companies A
and B are the sellers.

4.7.3.4 Income approach and associated methodologies


The income approach and its associated valuation methodologies provide the most
accurate results for valuation purposes and should be applied in all cases where a Life
of Mine Plan (LoM Plan) has been drafted, or where all the information that is used
to derive a LoM Plan is available. In order to define a JORC compliant reserve, the
appropriate information will be available.
The income approach together with its valuation methodologies is based on the calcu-
lation of FCF for the operation followed by the determination of the NPV of those future
cash flows. The NPV is determined using the DCF methodology, or a variation of this.
For a mining operation, the key input parameters that will support the use of a DCF
NPV will include:

• Statements of Reserves and Resources (at stipulated cut-off grades)


• Mining production profiles over selected, regular periods
• In-situ grades with anticipated recoveries or yields
• Working cost assumptions and forecasts, preferably on a unit of commodity basis
($/oz, $/lb, $/t, etc.)
• Anticipated capital expenditure forecasts
• Any other factor that may influence the fundamental value (working capital
changes, taxation and royalty policies, profit or income participations, free equity
carry or preferential participations, environmental considerations, potential
retrenchment liabilities and future closure-related costs)

In addition, the specific economic parameters that may be required include:

• The mineral’s projected price (existing spot price, forecast spot price and hedged
price). Co-products and by-products are included here.
• Exchange rate assumptions for products that generate revenues in a foreign currency.
Green and sustainable mining 413

• Interpretation and application of royalty, taxation and other cash-flow impacting


issues.
• The calculated discount rate that is to be applied to the resulting cash flows.

4.7.4 The time value of money


Due to inflation over time, all monetary figures, including values, must account for
historical and future inflation and escalation. This is referred to as “nominal terms”
cash flows, while in the absence of inflation or escalation, the term to describe finan-
cial figures today or at a specific point in time is “real terms”.
This introduces the concept of the time value of money, which is a crucial compo-
nent of any valuation methodology that looks at historical or future figures in order to
calculate a value reflected in today’s terms (money-of-the-day terms). Inflation, esca-
lation, deflation and de-escalation are all typically incorporated into a DCF NPV
analysis.
The DCF valuation method incorporates the modelling of key input data to
determine a final free cash flow (FCF) for each forecast period (e.g., each year).
The method considers that for any initial investment in a mining project, the inves-
tor will derive a minimum return on that investment through the future cash flows
of that project or investment. This minimum return will be at least equal to the
investor’s hurdle rate (a return greater than the investors actual cost of funds). The
hurdle rate represents the minimum return of a project below which the decision to
invest or develop a new project will be negative, and above which the project may
be developed. The key input data referred to above will typically be incorporated in
an LoM Plan, such as that accompanying a pre-feasibility or a bankable feasibility
study, or being associated with an existing mining operation looking to expand or
recapitalize.
Working capital balances, which are included in a DCF modelling exercise, are gen-
erally based upon historical creditor and debtor balance ratios, which are then applied
to forecast revenues and working costs to determine future working capital balances.
The NPV is then calculated from the resultant cash flow, with the NPV being calcu-
lated at a pre-determined discount rate or, more accurately, at a variable or dynamic
discount rate (a discount rate that typically differs from one period to the next). The
calculated NPV and associated ROI, which may be measured by the project’s internal
rate of return (IRR), are looked at together to determine the operation’s implicit value
and its likely ROI (economic or investment attractiveness).
The NPV of the project, derived from the DCF method, is denoted by:

 n 
NPV =  ∑
 t = 0
C f /(1 + r)t − I c 


Where:

Cf = forecast FCF at time t in real or nominal terms


t = time or period in which the cash flow occurs
r r = real or nominal discount rate
Ic = present value of the initial development capital expenditure
414 Green and sustainable mining

The DCF method of valuation does not adequately incorporate management


inflexibility. This then suggests that DCF valuations may give rise to projects being
inaccurately valued due to management being able to decrease downside risks whilst
simultaneously increase the upside potential of a project, at many decision points
along that operation’s life cycle.
In addition, as far as DCF methods are concerned, the inability to defer investment
decisions could introduce high opportunity costs that cannot be factored into a DCF
analysis. One of the methods that can be used to address these issues is through using
real options analysis or the option pricing method, which will not be discussed in this
book.
As can be expected, the technical inputs of the project will not vary from one anal-
ysis to another. The more comprehensive and reliable the technical forecasts, with
greater accuracy accompanying a definitive/bankable feasibility study, the greater the
certainty that can be placed upon the results of the valuation. However, as much as
significant effort and cost go into developing a feasibility study, it is unfortunate that
the same cannot be said about discount rates.

• Discount Rates – Valuers base the NPV discount rate on the corporate cost
of capital, necessary to calculate the present-day value of the operation. The
cost of capital reflects the total cost that will be incurred in order to fund an
acquisition, development, recapitalization or other need associated with the
operation. The cost of capital is calculated as the weighted cost of the vari-
ous sources of funding, typically equity, debt and, where available, preference
shares.
The weighted funding cost is known as the weighted average cost of capital
(WACC), which is the calculated rate at which forecast cash flows must be dis-
counted such that the capital raised by the company generates a return (cash
flow return) at least equal to the total funding cost. Companies typically tar-
get investment returns in excess of their WACC. Assuming the WACC does not
change over time, then the WACC should be the discount rate. However, as is
often the case, if the WACC is altered by including additional risk and uncer-
tainty factors such as pure project risk or an additional hurdle rate, then the
calculated WACC will not be the same as the discount rate. Logically, in this
case, the discount rate will be higher than the WACC.
• Weighted Average Cost of Capital (WACC) Determination – Since WACC is deter-
mined as the weighting of the cost of debt and the cost of equity (CoE), a valuer
needs to calculate the cost of debt and the CoE and then attribute the relative
weightings to each of them.
An applicable discount rate can be determined for a specific operation, based
on industry-derived factors, including an operator’s calculated IRR, non-tech-
nical risk factors impacting the IRR and possible project risks. Economic and
finance theory offers tools to calculate discount rates, calculating the WACC
through determining the CoE by using the capital asset pricing model (CAPM)
and obtaining debt funding rates. The WACC represents the weighted average
cost of funds on offer, and is based on the following:

rWACC = re p e + rd pd + rp p p
Green and sustainable mining 415

Where:

rWACC = weighted average cost of capital


re,d,p = proportional costs of equity, debt and preferred stock (%)
pe,d,p = percentage of equity, debt and preferred stock, the sum being
1.00 (i.e., pe + pd + pp = 1.00)

The CAPM formula, to calculate the CoE, is designed to generate the return equal
at least to the return that should be obtained in the event when the investor places
the same capital in an alternative investment demonstrating the same level of risk.
If an operation is funded solely using equity (i.e., no debt), then only the CoE is
to be considered, which means that the CAPM result will equal the WACC. The
CAPM is determined as follows:

re = f + Rβ

Where:

re = expected return on the stock


f = risk-free return (country or sovereign risk)
R = risk premium of market returns over long-term risk-free rates
(typically 5–7%)
β = Beta factor for the stock (coefficient of systematic risk)

So therefore:

rWACC = p e (f + Rβ ) + rd pd + rp p p

Either pre-tax or post-tax calculations must be conducted. That is, if the cash flows
being discounted are on a pre-tax basis, then the WACC and CAPM models must be
performed on a pre-tax basis. Therefore, the contributing components to the determi-
nation of the cost of capital must be on a pre-tax basis too, such as the risk-free rate
and the costs of debt, equity and preference shares. The reciprocal applies to a post-
tax position. Most DCF NPVs are performed on post-tax cash flows, so the WACC
must also be a post-tax figure.

4.7.5 Market risk (a)


The total market risk is the volatility of the share price histories of all listed stocks on
the securities exchange. Typically three-month, six-month and one-year volatilities are
determined. The volatility to be used in the CAPM is generally calculated over the long-
est period, and must exclude periods (influences) of extreme or event-specific volatilities.
The Beta coefficient of a particular asset is a measure of the systematic risk relating
to that asset. This systematic risk is the portion of total risk that cannot be eliminated
or diversified away by holding a fully diversified investment portfolio. A Beta coeffi-
cient reflects the sensitivity of an asset’s value to economic variables that affect the
values of all risky assets, including economic growth rates, exchange rates, interest
rates and even inflation rates.
416 Green and sustainable mining

A Beta coefficient of greater than 1 means that the asset hosts more systematic
risk than the market portfolio of risky assets. A Beta coefficient of less than 1 implies
the asset conveys less risk than the whole market. A Beta coefficient of exactly one is
impossible.
In the case of an unlisted asset such as an exploration concession, its Beta coeffi-
cient can be determined by manipulating the Beta coefficient of comparable listed
companies. This makes use of a proxy Beta.
A market-based Beta is not perfect. The issues include:

• Betas vary over time.


• Betas reflect the volatility of a share price and not of a specific asset within a company.
• Betas vary as the market varies and not independently of it.
• Relative Betas will vary so that a perfect correlation is improbable.

Furthermore, the assumptions incorporated in the CAPM include issues as follows:

• Seldom do investors emulate each other’s expectations of risk and return.


• Investors may not make investment choices on the basis of risk versus return.
• Returns being measured by the mean, with risks being measured by the variance.

In the worked example below, it is assumed that the proxy Beta applies to a listed company
that carries a specific level of debt. Therefore, the obtained geared or re-levered Beta has
been calculated from the ungeared or unlevered Beta based on the debt-to-capital ratio
and the applicable taxation rate. Therefore, the following applies:
Assuming a proxy Beta (i.e., a geared Beta from a proxy company) and different
tax-paying rates:

Unlevered Beta = β geared /((1 + pd /p e )(1 − tax rate))

and:

Regeared Beta = β unlevered × ((1 + pd /p e )(1 − tax rate))

By way of an example, consider the following (no tax difference as the example consid-
ers an asset in the same country as the proxy), say:

β = 1.48 (a geared, proxy Beta of a company with structure of $100m


capital and debt of $40m)

then:

Ungeared β = 1.48/(1 + 40/100) × (1 − 30%) = 1.16

and if the targeted gearing is a debt:capital ratio of 50%, then:

Regeared β = 1.16 × (1 + 0.5) × (1 − 30%) = 1.56


Green and sustainable mining 417

Overall, the WACC denotes a one-period pricing model solution, with a one-period risk
premium. This single period representation of risk or uncertainty tends to be inappro-
priate for mineral projects for many reasons, including that mineral projects often have
lives over many periods of varying growth and decline, commensurate with continuously
varying debt and equity levels. Over these periodic variances, WACC will vary due to:

• The term structure of the modelled interest rates (varies off the base rate)
• The variable cost of risk
• Country-specific macro-economic influencing factors (e.g., sovereign risk, CPI)
• Specific aspects of project risk and uncertainty
• Socio-political and ESG factors varying over time

4.7.6 Country or sovereign risk


Sovereign risk presents a factor that incorporates civil stability or unrest, economic
policies, a country’s leadership and their principles and other factors. The compo-
nents of sovereign risk are categorized in Table 4.10.
The long-term risk-free interest rate, which is a reflection of the economics asso-
ciated with sovereign risk and is provided as a return or yield on a government-is-
sued instrument such as a bond, is obtainable from numerous sources. These data
are quoted as a pre-tax yield on government bonds, being the guaranteed, pre-tax,
risk-free return or yield on money invested in that specific bond.

Table 4.10 Components of sovereign risk

Risk Category Specific Risk Component

Political Risk Government stability


Political parties
Constitutional risk
Quality of government
Foreign ownership policy
Foreign policy
Government crisis
Taxation instability
Environmental policy/protectionism
Land claims and protected areas
Geographic Risk Transportation
Climate
Economic Risk Currency stability
Foreign exchange restrictions
Social Risk Distribution of wealth
Ethnic or religious differences within the indigenous population
Corruption
Labour relations

Bond yields (the risk-free rate of return) depend on the bond’s time to maturity.
They are quoted in nominal and pre-tax money terms and therefore the CPI rate will
418 Green and sustainable mining

Table 4.11 Variable (dynamic) discount rates (example)

Parameter Units Start Peak Steady state

Debt Level US ’mill 800 1,200 100


Capital Level US$ ’mill 200 200 200
Debt/Capital Ratio % 80% 86% 33%
Effective Tax Rate (CIT) % 0.00% 29.00% 37.88%
Borrowing Cost (LIBOR+bp) % 7.14% 7.14% 7.14%
Unlevered Beta 0.85 0.85 0.85
Equity Risk Premium % 6.00% 6.00% 6.00%
Risk-free Rate % 5.12% 5.12% 5.12%
Re-levered Beta 4.25 4.47 1.11
CoE % 30.62% 31.94% 11.80%
Proportional CoE % 6.12% 4.56% 7.87%
Proportional Cost of Debt % 5.71% 6.12% 2.38%
WACC nominal % 11.83% 10.68% 10.25%
WACC real (at 5.5% CPI) % 6.00% 4.91% 4.50%

have to be deducted from the returns to obtain the risk-free rate in real money terms.
Naturally, if post-tax (vs. pre-tax) cash flows are discounted, the yields must be dis-
counted at the post-tax (vs. pre-tax) cost of capital.
• Calculating the CoE (CAPM) – The calculation of the CoE follows (refer to Table
4.11 for the determination of the Beta used below, and assume a tax rate of 30%
and a risk-free rate of 2.40%):
To calculate the CoE from the above β example, with R at 5%:

re = f + Rβ
= 2.40 + (0.05 × 1.56)
= 10.20% (pre-tax)

and:

= 2.40 × (1 − 30%) + (0.05 × (1 − 30%) × 1.56)


= 7.14% (post-tax)

As stated previously, the WACC will vary over time due to debt:capital and equity:-
capital ratios varying over time. Table 4.11 demonstrates the impact of varying levels
of debt and equity over the life of the asset and the impact that this variation has on
the calculated WACC.
The fact that WACC varies over time is discussed in detail in Lilford et al. (2018).

4.7.7 Project risk


The inherent risks and uncertainties associated with a project are never constant and
are never known for all periods. They will change to variations in the sources and
Green and sustainable mining 419

magnitudes of the influences underlying the mineral project, which will be evidenced
by a changing project profile. This profile change will be noted as:

• Varying grades
• Varying production rates
• Changing unit cost profiles
• New project developments
• Dynamic exploration rates
• Primary development rates
• Hedging philosophies
• Various other lesser alterations (recoveries, dilution, efficiencies, expansions and
partial closures)

These risks and uncertainties should not be factored into a WACC, but should rather
be kept separate and treated through the use of real options, sensitivity analyses, etc.
If, however, the valuer is adamant that these risks and uncertainties are to be included
in a discount rate, they can be separated from the WACC to show the explicit impact
of these variations to WACC.
Table 4.12 shows a summary of the key input determinants that impact the final
cash flows, and then also shows the resultant NPVs based on a single WACC ($585
million), on a dynamic WACC ($834 million) and on a dynamic WACC with project
risk and uncertainty ($677 million).

4.7.8 Valuing shares (scrip)


Scrip is another word for shares. There are two forms of shares, being listed (stock
exchange and therefor public) and unlisted (privately held).
Listed shares are simple to evaluate since the value of those shares is simply the
listed market price ($ or c per share) multiplied by the number of shares held or being
transferred. It is good practice to not use the spot price of the shares (a spot price is the
share price at the latest point in time, which often is not reflective of a period’s price,
and also may be chosen to represent a low price point, reducing the taxable transfer
value), but to rather apply a volume weighted average price (VWAP) over an appropri-
ate number of days preceding the property and shares transfer.
The VWAP of a listed company is calculated as the volume of shares traded over
a trading period (day) multiplied by the actual share price of each trade on that day.
This is conducted over several days, with the resulting VWAP being the sum of the
figures divided by the total number of shares traded over that period.
Example: VWAP. A listed company has had its shares traded over a 10-day period
as shown in Table 4.13. The average share price of each day’s trade is also shown. The
VWAP is then calculated by multiplying the daily volume by the daily price, adding
all of the products together, and dividing that sum by the sum of the volumes alone.

4.7.9 Mining financing options


From the discussion around determining the cost of funding, it is apparent that there are
two main sources of mine financing: equity and debt. However, increasing sophistication
over time has introduced a myriad of hybrid funding instruments, including convertible
notes, warrants, metal streams, offtake contracts and many others.
420 Green and sustainable mining
Table 4.12 A typical DCF showing NPVs at varying discount rates

Summary Year 0 1 2 3 4 5 6 7 8 9 10 35

Revenue $’000 67,106 163,360 202,696 216,796 222,216 227,772 233,466 239,303 245,285 251,417 116,528
Costs $’000 38,800 96,815 120,127 128,484 131,696 134,988 138,363 141,822 145,368 149,002 69,060
Operating profit $’000 28,306 66,545 82,569 88,312 90,520 92,783 95,103 97,480 99,917 102,415 47,468
Capex $’000 - 189,113 96,920 33,114 6,424 6,585 6,749 6,918 7,091 7,268 7,450 -
Tax (nominal case) $’000 8,492 10,602 11,684 11,3334 13,733 14,128 14,487 14,853 15,224 22,702 12,114
Free Cash Flow $’000 - –169,299 –40,977 37,770 70,555 70,203 71,906 73,697 75,537 77,425 72,263 35,355
(FCF)
FCF real terms $’000 - - –165,169 –39,003 35,073 63,919 62,049 62,004 61,999 61,996 61,996 56,452 14,897

VALUES 0% 1,585,465 $’000 22.87% IRR

5.20% 585,111 $’000


Dynamic DR 833,956 $’000
Dynamic DR+risk 677,141 $’000

(Continued)
Table 4.12 A typical DCF showing NPVs at varying discount rates (Continued)

Differential (variable)
Discounting 0 1 2 3 4 5 6 7 8 9 10 35

Discount rate % 5.20% 5.20% 5.20% 6.98% 7.54% 4.21% 3.87% 3.55% 3.25% 2.97% 2.72% 1.37%
Cashflow $’000 - –165,169 –39,003 35,073 63,919 62,049 62,004 61,999 61,996 61,996 56,452 14,897
Discount factor 1 1.000 0.951 0.904 0.845 0.785 0.754 0.725 0.701 0.679 0.659 0.642 0.440
(rate)
Discounted cash $’000 - –157,000 –35,240 29,622 50,197 46,758 44,982 43,435 42,065 40,853 36,214 6,552
flow
Progressive $’000 - –157,000 –192,240 –162,618 –112,421 –65,663 –20,681 –22,753 64,819 105,672 141,886 833,956
discounted value
Pure Project Risk % 0.0% 0.0% 0.1% 0.2% 0.3% 0.4% 0.5% 0.6% 0.7% 0.8% 0.9% 3.4%
Discount Rate
Project Risk 1 1.000 1.000 0.999 0.9997 0.994 0.990 0.985 0.979 0.972 0.965 0.956 0.555
Discount Rate
Risk-adjusted value $’000 - –157,000 –35,205 29,533 49,897 46,293 44,313 42,534 40,907 39,413 34,626 3,638
Progressive Value $’000 - –157,000 –192,205 –162,672 –112,775 –66,481 –22,168 20,366 61,273 100,686 135,311 677,141

Green and sustainable mining 421


(risk adjusted)
Progressive IRR on % –64.24% -25.46% -7.62% 2.58% 8.88% 12.99% 15.78% 17.57% 22.87%
Cashflow

Differential discounting value Incl project risk 677,141 A$’000 22.87% IRR (check)
422 Green and sustainable mining

Table 4.13 VWAP calculation

Date Shares traded Trade price Volume price


No. cps $

Wednesday 154,000 76 117,040


Thursday 598,020 79 472,436
Friday 454,927 82 373,040
Monday 76,458 78 59,637
Tuesday 975,234 81 789,940
Wednesday 505,213 83 419,327
Thursday 367,243 84 308,484
Friday 287,354 86 247,124
Monday 595,275 84 500,031
Tuesday 486,391 84 408,568
Total 4,500,115 3,695,628
VWAP 82.12

Mining finance is required to achieve the following at different stages of the mining
life cycle:

• To cover research and/or exploration costs [typically equity only, including pri-
vate equity, seed capital from High Net Worth individuals (HNWs)].
• To complete Scoping Studies (SS), Pre-Feasibility Studies (PFS) and to complete
Bankable/Definitive Feasibility Studies(B/DFS) (typically equity only).
• To deliver Front End Engineering Designs (FEED), Build-Own-Operate contracts
(BOO), Build-Own-Operate and Transfer contracts (BOOT), etc. (pseudo-debt,
potentially with convertible or hybrid instruments).
• To construct and commission a project (equity, debt, convertible, hybrid
instruments).
• Ramp up production (debt, equity, convertible, hybrid instruments).
• Recapitalize, expand, further develop a project (cashflow, debt, convertible, hybrid
instruments).

4.7.10 Equity
With the development of any mining project, the first money to be spent on the project
(exploration and initial project development) will be equity capital.
From the very early stages of a project’s life, the only funding available to be used
for exploration activities is seed capital. This is equity funding provided by the found-
ers of the company through which the project is owned. This seed equity is unlisted, is
very high risk, and tends to be a small amount (possibly a few tens of thousands up to
potentially a hundred thousand dollars or two).
Once the project has advanced sufficiently with the limited funds, high networth
individuals (HNW) will typically be approached to be offered an opportunity to add
their own equity capital into the speculative business for a pro-rata equity stake, again
with the company not being listed on an exchange.
Private equity groups may also be approached for funding at this stage.
Only once larger amounts of capital are required (greater than a few million dol-
lars), the founders (seed investors) may take the company through an initial public
Green and sustainable mining 423

offering (IPO) and list on an appropriate securities exchange. Once listed, a number
of new sources of funding becomes available, including:

• Rights offers
• Private placements
• Convertible (equity) instruments
• Warrants
• Joint ventures (JV) (although this may be done without listed equity)
• Other sources, where equity is used as security to obtain the requisite funding

The advantages of equity financing for a mining venture are:

• The ability to fund higher risk projects.


• No obligation to pay a dividend on ordinary shares (contrast this with interest on debt).
• No maturity date for repayment of initial capital contribution by shareholders.
• Shareholders’ liability is limited to residual assets after the claims of all other parties
have been first met, but it does not extend to the shareholders’ other personal assets.
• Venture capital aims for high capital gains by disposal of investment rather than
waiting for dividends.

The disadvantages of equity financing for a mining venture are:

• There are dilution effects from additional equity issues (as opposed to debt finance
which does not cause dilution of ownership).
• Variable availability of funds as a function of market sentiment and commodity
price cycles.
• The complexity and high transaction costs of an equity issue.

4.7.11 Debt
Debt funding is (almost always) cheaper than equity funding, as demonstrated previ-
ously in the calculations of Weighted Average Cost of Capital (WACC).
A provider of debt finance, such as a bank, will apply its own market-based eco-
nomic parameters (commodity price, exchange rate, etc.) to determine the cash flows
of the project, and will also include certain covenant ratios to the free cash flows
(FCF) to gauge the robustness associated with debt serviceability that is likely to be
generated by the proposed operation.
Debt may be provided by commercial or merchant banks, at lending rates that are
benchmarked off a base funding rate. The base funding rate will be the rate at which that
country’s central bank lends to the commercial or merchant banks, and include funding
rates such as London Inter Bank Offer Rate (LIBOR), Australia’s Bank Bill Swap Rate
(BBSW), South Africa’s Johannesburg Inter Bank Acceptance Rate (JIBAR), etc.
A few facts surrounding the debt market are summarized below:

• Debt amounts may be syndicated (a primary lender syndicates a portion of the


debt to other lenders). If debt is syndicated (a “syndicated loan”), a lead bank is
appointed and the borrower only deals with the lead bank – i.e., the lead bank will
syndicate the loan. This is not the same as mezzanine/subordinated debt.
424 Green and sustainable mining

• Debt may come with mezzanine (subordinated) debt, typically being more expen-
sive debt due to attaching warrants arising from the necessary additional lending
rates. This is due to the mezzanine lender having no security on the mezzanine
facility – i.e., the security is tied up with a primary lender already.
• Mezzanine debt (subordinated security) is second-ranking (subordinated rights
to security, so has more to lose in the event of default), but comes with embedded
convertible options or warrants.
• Debt may come with a moratorium period, where capital and interest are not
repaid for a fixed period (during ramp-up, by way of example) while interest is
capitalized (accrues), or a capital-only moratorium period where no capital is
repaid but the interest is serviced over that initial moratorium period. Typical
moratorium periods are between 12 months and 24 months.
• Financial institutions only provide debt if adequate equity is already invested or
is available to be invested.
• Equity is always drawn down first (this is a bank/lender’s requirement).
• Typically (and dependent on the borrower), the debt:equity ratio may be around
40:60 (i.e., 40% debt and 60% equity).
• Debt tenures are often seldom greater than 6–8 years (project life permitting).
• Lending rates are based on the bank’s own borrowing rate (e.g., LIBOR, BBSW,
JIBAR, etc.) plus an additional rate.
• The additional rate or margin is added to the base rate (e.g., LIBOR plus “x” basis
points).
• A basis point is a pip (percentage in point), being 0.01%: so 150bp is 150 pips,
which is 1.5%. So if LIBOR is, say, 1.95%, then a borrowing rate of 250bp above
LIBOR means the borrower’s cost is 4.45% per annum, compounded daily or
monthly.

Lenders or debt providers require that certain covenants and conditions are met to
provide comfort that the borrower is not approaching a default position. The impor-
tant covenants that a lending bank will impose on a borrower include:

• A debt service cover ratio (DSCR) – the ability to repay debt with comfort:
• DSCR = (Net Operating Income)/Total Debt Service
Where:
− Net Operating Income = Revenue − (Certain Operating Expenses)
− Total Debt Service = Current Debt Obligations
− Total Debt Service = [Interest × (1 − Tax Rate)] + Principal
• A loan life cover ratio (LLCR) – a solvency test:
• The LLCR can be calculated using a shortcut: dividing the NPV of project
FCF by the present value of the debt outstanding. In this calculation, the
weighted average cost of debt is the discount rate for the NPV calculation and
the project “cash flows” are more specifically the cash flows available for debt
service [cash flow after debt service (CFADS)].
• A cash sweep on excess earnings (up to a 30% cash sweep)
• Hedging (up to 70% of annual production over life of loan)
Green and sustainable mining 425

To secure debt funding, a mining company should have existing positive net cash
flows or expect to become cash-flow-positive in a sustainable manner in and over the
foreseeable future.
The advantages of debt financing for a mining venture are:

• Flexible availability of funds which can be used when actually needed


• Lower cost of funds and transaction costs compared with other financing options
• The tax deductibility of interest expenses which leverages the returns on share-
holders’ equity
• No dilution of ownership (contrast that with share issues which cause dilution of
shareholders’ equity)
• Matching loan maturities to intended users

The disadvantages of debt financing for a mining venture are:

• Larger mining companies have greater choice and ease of debt instruments than
smaller ones.
• The higher level of debt brings about more “financial risk”.
• Floating security over company’s assets extend financial risk across to other
projects.
• Restrictive covenants in mine project finance that may limit the capacity to use
other company assets as security to raise funds for other projects.
• Interest expenses are unrelated to fluctuating profit levels on the mine.

References
Bartels, R. (2019) Mining new value from the circular economy. Accenture Research Report.
April 16.
Benckert, A. & Eurenius, J. (2001) Tailings dam constructions. Seminar on safe tailings dam con-
structions. Gallivare, Swedish Mining Association, Natur Vards Verket, European Commission.
pp. 30–36.
BHP (2019) ESG Briefing: Tailings Dams (Escondida). June.
Brereton, D., Joyce, S. & Kemp, D. (2020) Tailings facilities and social performance: Why commu-
nities matter. Social & Environment webinar – AusIMM. November.
CIMVal (2019) Canadian Institute of Mining, Metallurgy and Petroleum. https://www.cim.org/
news/2019/2019-cimval-code-for-the-valuation-of-mineral-properties-adopted-by-cim/.
Coloradoans Against Resource Destruction (2008) Colorado Company (colorado-corp.com)
http://colorado-corp.com/co/coloradoans-against-resource-destruction.
Colwell, B. (2017) Sustainability & the Circular Economy Finally Break into the Mining Industry.
Blog. November 2.
Cooney, J. (2016) Social Licence to Operate: Revisiting the Concept. Presentation to Ryerson
University. June 28.
Corder, G. (2021) The Circular Economy: A Sustainable Future for Mining and the World.
Sustainable Mineral Institute, University of Queensland.
Engels, J.M. & Dixon-Hardy, D.W. (2007) Guidelines and recommendations for the safe operation
of tailings management facilities. Environ Eng Sci. Vol. 24, No. 5, pp. 625–637.
Environmental Law Alliance Worldwide (2010) Guidebook for Evaluating Mining Project EIAs.
July. www.elaw.org/files/mining-eia-guidebook/Full-Guidebook.pdf.
Frerot, A. (2018) The Post, Planet. Veolia. October. www.veolia.co.uk/sites/g/files/dvc1681/files/
document/2018/10/Veolia%20UK%20_%20Planet%20Magazine%2016.pdf.
426 Green and sustainable mining

Frimpong, S. (1992) Evaluation of mineral ventures using modern financial methods. Unpublished
Doctoral Dissertation. University of Alberta. 1982.
GRID-Arendal (2020) Global tailings portal project. Norway. January. http://tailing.grida.no/.
Harraz, H.Z. (2014) Review of JORC Code & Mining Public Records presentation. August 5.
Hartmann, H.L. & Mutmansky, J.M. (2002) Introductory Mining Engineering. John Wiley &
Sons, New Jersey, US.
Husseini, T. (2018) The future of mining: Eight bold industry predictions. Mining Technology.
September 26.
International Commission on Large Dams (ICOLD) & United Nations Environment Programme
(UNEP) (2021) Tailings dams risk of dangerous occurrences—Lessons learnt from practical
experiences. Bulletin 121. March.
International Council on Mining & Metals (ICMM) (2012) Community development toolkit.
https://commdev.org/pdf/publications/ICMM-Community-Development-Toolkit.pdf.
International Council on Mining & Metals (2020) Global tailings report: New Global Industry
Standard on Tailings Management aims to improve the safety of tailings facilities in the mining
industry. August.
International Institute for the Environment and Development (2002) Mining, Minerals and
Sustainable Development Project.
International Standard Organization (2010) ISO 26000 Guidance on Social Responsibility.
Jain, P.K. (2020) An introduction to mineral economics & the role of the geologist. e-Training
on: “Basic course on Resource Estimation for Mineral Deposits”. Indian Bureau of Mines. June.
Joyce, S. & Thomson, I. (2000) Earning a social licence to operate: Social acceptability and
resource development in Latin America. CIM Bull. Vol. 93, No. 1037, pp. 49–53.
Kirkey, J. (2014) Eco-friendly mining trends for 2014. Natural Resources Canada.
Kuchling, K.J. (2019) Sustainable mining: What is it really? K.J. Kuchling Consulting Ltd blog.
December 6.
Lacy, P. & Rutqvist, J. (2015) Waste to Wealth: The Circular Economy Advantage. Kindle Edition.
Lilford, E.V. (2004) Advanced considerations, applications and methodologies in the valuation
of mineral properties. Doctoral thesis submitted to the Faculty of Engineering and the Built
Environment, University of the Witwatersrand, Johannesburg.
Lilford, E.V. (2012) Advanced methodologies for mineral project valuation. Australian Institute
of Geoscientists. Bulletin No. 53. ISBN: 1 876118 41 5. ISSN: 0812 60 89.
Lilford, E.V., Maybee, B. & Packey, D. (2018) Cost of capital and discount rates in cash flow val-
uations for resources projects. Resour Policy. Vol. 59, pp. 525–531.
Manifesto for a Resource-Efficient Europe (2012) European Resource Efficiency Platform, European
Commission. December.
Martin, T. E. & McRoberts, E. (1999) Some Considerations in the Stability Analysis of Upstream
Tailings Dams. Amec Foster Wheeler. January.
Matthews, E. & Spearing, A.J.S. (1999) Tailings disposal practice in South Africa. Paste Technology
for Thickened Tailings. University of Alberta, Edmonton, Canada. November.
McDowall, W., Geng, Y., Huang, B., Barteková, E., Bleischwitz, R., Türkeli, S., Kemp, R. &
Doménech, T. (2017) Circular economy policies in China and Europe. J Ind Ecol.
McMullen, L. (2007) Risk Assessment Management (RAM). Mining Safety and Health Administration
(MSHA).
Medkova, K. (2016) Urban mines – The mines of circular economy. Smart Cities in Smart Regions
Conference.
Meech, J. A. (n.d.) APSC 150 – Case study 3: Sustainable mining; is that possible? Norman Keevil
Inst. Min. Eng, UBC & Centre for Environmental Research in Minerals, Metals and Materials.
https://slideum.com/doc/1206094/case-study-3---introduction-to-sustainable-mining
Meilani (2018) Barriers to the circular economy: The blocking effects of regulations. Blog. Material
Trader. December 8.
MIT Mission (2016) http://web.mit.edu/12.000/www/m2016/finalwebsite/index.html.
Green and sustainable mining 427

Mkandawire, M. & Oakes, K. (2015) Mending mining. Canada’s map to sustainability. Altern J.
March. www.alternativesjournal.ca/climate-change/mending-mining/.
Nicholson, R.V., Ludgate, I., Clyde, E. & Venhuis, M. (2012) The successful reclamation of acid
generating tailings in the Elliot Lake Uranium District of Canada. 9th International Conference
on Acid Rock Drainage (ICARD), Ottawa, Canada.
Oberle, B. (2020) Global tailings review. International Council of Mining & Metals, UN
Environmental Programme & Principles for Responsible Investment. August 5.
Ostensson, O. & Roe, A. (2017) Sustainable Mining. International Labour Organization.
Owen, J.R., Kemp, D., Lèbre, E., Svobodova, K. & Pérez Murillo, G. (2020) Catastrophic tailings
dam failures and disaster risk disclosure. Int J Disaster Risk Reduct. Vol. 42, p. 101361.
PriceWaterhouseCooper (2019) Mine 2019: Resourcing the Future. www.pwc.com/gx/en/ener-
gy-utilities-mining/publications/pdf/mine-report-2019.pdf.
Rau, S. (2019) Options for Urban Mining and Integration with a Potential Green Circular Economy
in the People’s Republic of China. Asian Development Bank. December.
Rhodes, A. (2020) A brief summary of the long history of risk management. Blog. Ventiv Technology.
Rio Tinto (2021) Smart mining. https://www.riotinto.com/en/about/innovation/smart-mining.
Robinsky, E.I. (1999) Thickened Tailings Disposal in the Mining Industry. E.I. Robinsky Associates
Limited, Toronto, Canada.
Schwab, K. (2017) The Fourth Industrial Revolution: What it means, how to respond. World
Economic Forum. June 29.
Society of Mining Professors (SOMP) (2019) Mines of the future. Version 1: Major findings
report. ISBN: 978-0-7334-3865-3.
Sonter, L.J., Ali, S.H. & Watson, J.E.M. (2018) Mining and Biodiversity: Key Issues and Research
Needs in Conservation Science. The Royal Society Publishing. https://doi.org/10.1098/
rspb.2018.1926.
Stephenson, P. (2017) Valuation of Mineral Exploration Properties. AMC Consultants.
The JORC Code (2012) Prepared by the Joint Ore Reserves Committee of The Australasian
Institute of Mining and Metallurgy, Australian Institute of Geoscientists and Minerals Council
of Australia.
Tiseo, I. (2020) Plastic waste worldwide: Statistics and facts. Energy & Environment – Statista.
September 4.
Topf, A. (2011) Company proposes to mine Arizona copper deposit using in-situ method.
December 22. Retrieved from: http://www.metalaugmentor.com/reviews/company-proposes-
to-mine-arizona-copper-deposit-using-in-situ-method/.
Tyson, R. (2018) Corporate Social Responsibility. Mining International Ltd. https://miningir.com/
podcast-the-role-corporate-social-responsibility-plays-in-a-successful-mining-company/.
Ulmer-Scholle, D.S. (2022) Uranium: How is it mined? Feb 22. Retrieved from: http://geoinfo.
nmt.edu/resources/uranium/mining.html.
Vick, S. (1990) Planning, Design and Analysis of Tailings Dams. BiTech Publishers. Richmond, BC,
Canada.
Vietti, A. (2020) Tailings dewatering trends over the last 40 years. Webinar for SAIMM Johannesburg
Branch. November 19.
Wikipedia (2019) Brumadhino dam disaster. www.en.wikipedia.org/wiki/Brumadinho_dam_disaster
World Economic Forum (2019) 7 trends shaping the future of mining and metals industry.
Nicolas Maennling and Perrine Toledano are co-curators of the World Economic Forum’s
Transformation Map on Mining and Metals. March.
World Mining Data (2020) International Organizing Committee for the World Mining Congresses.
Federal Ministry of Agriculture, Regions & Tourism, Austria.
Index

Adits 225–226 Calder, P. et al. 29–30


Agioutantis, Z., Barley,D., Karmis, M. & Call, R.D. 52–52
Elrick, S. 174 Camus, J. 36
Akande, J.M. & Saliu, M.A. 112–114 Carver, L. & Rauch, H.W. 176
Alehossein, H. & Poulsen, B.A. 104–105 Cave Mining 246–252
Alimak Raising 224–225 Chain pillars 114–117
Analytical methods 151–153 Chamber of Mines Research Organization 242
Anderson, N.L., Apel, D.B., Ismail, A., Chevron blasting 26
Kovin, O. & Dzelic, V. 123 CIM Val 404
Angle-of-Draw 123–124 Circular mining economy 351–357
Anthracite 61–63 Clark, C. & Molinda, G.M. 69
Area mining 13–15 Coal bump precautions 160–161
Armoured Face Conveyor 106–107 Coal classification 58–63
Artificially supported mining methods 262–281 Coal exploration 64
Atlas Copco. 223, 243, 247, 254, 257, 265, 266 Coal formation 56–57
Coal mine access and development 80–91
Backfill 334–339 Coal mine hazards, safety and risk 161–165,
Barczak, T.M. 155 Coal mine planning 74–79
Bartels, R. 356–357 Coal mine planning and design 66–67
Barton, N. 200–206 Coal mine rehabilitation 165–172
Barton, N., Lien, R. & Lunde, J. 206–207 Coal mine rockmass classification 67–74
Beam building 144–145 Coal mine safety training 164
Beam theory 151–152 Coal mine waste rehabilitation 167–172
Belem, T. & Benzaazoua, M. 338 Coal mining subsidence 175–179
Bench blast 28 Coal rockburst precursors 180–181
Bench failure 34 Coal usage 58
Bench height and angle design 32–34 Colwell, B. 349
Benckert, A. & Eurenius, J. 369–370 Components of an explosive 283–284
Bhattacharya,S. & Singh, M.M. 177 Computer design 51–52
Bieniawski, Z.T. 46, 67, 153, 206–212, 311–312 Computer modelling method 153–155
Bituminous coal 60–61 Concepts of green mining 347–350
Blasting 37 Contour mining 20–21
Blasting monitoring 304–305 Controlled blasting techniques 300–301
Blasting principles 287–288 Conventional dam construction 367–370
Block Caving 246–252 Cook, N.G.W. & Salamon, M.D.G. 238–240
Booth, C.J., Spande, E.D., Pattee, J.D., Cooney, J. 363–364
Miller, J.D & Bertsch, L.P. 176–177 Corder, G. 355
Brady, B.H.G. & Brown, E.T. 198 Country risk 417–418
Bereton, D., Joyce, S. & Kemp, D. 374 Cut and fill mining 263–271
Brown, E.T. 234–235
Brown, E.T., Bray, J.W., Ladanyi, B. & Hoek, Dagdelen, K. 37
E. 154 Das, M.N. 92
Bruhn, R.W. & Speck, R.C. 178 Das Sharma, P. 288
Burn cut 301–302 Debt 423–425
Index 429

Deere, D. 199–200 Green mining 343–351


Deere, D.U. 46 Grimstad, E. & Barton, N. 209
Deere, D.U. & Deere, D.W. 200 Ground Reaction Curve 154–155
De La Vergne, J.N. 220
Denkhaus, H.G. 240 Hagan, T.N. & Mercer, J.K. 25,26
Design of longwall face supports 117–120 Haines, A., Voulgaris, P., Walker, D. &
Design Rock Mass Strength 211–212 de Bruyn, I. 44
De Souza, E. 256 Hamrin, H. 219, 227, 244, 246, 248, 249, 255,
Dominy, S.C., Noppe, M.A. & Annels, A.E. 217 260, 266, 277, 280
Diamond blasting 27 Hard rock exploitation 196–197
Diamond drilling 194–196 Hard rock exploration 190–196
Dragline 23 Hard rock geology 187–190
Drift and fill mining 262–263 Hard rock mine access and development
Drilling and blasting 24–27, 281–319 218–230
Drilling and blasting equipment 292–296 Hard rock mine design and planning
Drilling & wedging 18–19 197–217
Dust 37 Hard rock mine planning challenges 217
Hard rock production cycle 213–214
Earthjustice 168 Hard rock room and pillar mining 253–256
Echelon blasting 26 Hard rock support design 305–339
Einicke, G.A. 158 Harraz, H.Z. 402
Empirical design 46 Hartman, H.L. 35–36
Empirical method 146–151 Hartman, H.L. & Mutmansky, J.M. 28–29,
Engels, J.M. & Dixon-Hardy, D.W 367–370 346
Environmental Impact Assessment 192–193, Hassell, R., Villaescusa, E., de Vries, R. &
358–360 Player, J. 293
Environmental Protection Agency 169–172 Hedley, D.G.F. & Grant, F. 97
Equity 422–423 Hedrick, N. 320–321
Experiential design 46 Hobba, W.A. 177–178
Experiential method 146 Hoek, E. & Bray, J.W. 42
Exploration 41 Hoek, E. & Brown, E.T. 147–148, 310
Exploration stages 191–193 Hoek, E., Kaiser, P.K. & Bawden, W.F.
Explosives and accessories 296–305 148–150, 271
Explosive properties 288–292 Hoek, E. & Marinos,P. 45
Hoek, E. & Wood, D.F. 310
Fejes, A.J. 173 Hormazabal, E., Alavrez, R., Russo,
Feng, X., Hudson, G.A. & Tan, F. 153–154 A. & Acevedo, D. 251–252
Forbes, B., Vlachopoulos, N. & Hyett, A.J. 323 Huang, P., Spearing, A.J.S., Fu, J., Jessu,
Fourie, G.A. & Dohm, G.C. 40 K.V., Wang, Z. & Ning, P. 137–138
Fourie, G.A. & van Niekerk, D.J. 66–67 Hunt, S.R. 127–128
Fourth Industrial Revolution 350–351 Husseini, T. 392–395
Frerot, A. 380,382 Hustrulid, W.A. 125
Frimpong, S. 215, 404 Hustrulid, W.A. & Bullock, R.L. 247, 252,
Future mining technologies 392–398 261, 305
Hybrid method 155
Geological strength index 147–151 Hydraulic Backfilling 135
Geotechnical bench and slope design Hydrology 42
analysis techniques 52–53
Geotechnical design 39–54 Iannacchione, A.T., Mark, C., Repsher,
Geotechnical models 46–52 R.C., Tuchman, R.J. & Jones, C.C. 93–94
Geotechnical Strength Index 45 Ingram, D.K. 127
Gertsch, R.E. & Bullock, R.C. 264, 268, 270, International Commission on Large dams 369
273–274, 279, 281 International Council of Mining & Metals
Goldfields 260 361–362, 373
Global coal reserves 57 International Institute for Environment and
Gosh, S. 74 Development 166
Gray, R.E., Gamble, J.C., McLaren, R.J. & International Standard Organization 364
Rogers, D.J. 121, 134 Intersection support 155–159
430 Index

Jager, A.J. 320–321 Mine waste repurposing 376–377


Jain, P.K. 403 Mineral economics 402–413
Jakubee, J. & Laubscher, D.H. 46, 210 Mineral reserves 400–402
Jayaranjan, M.L.D., van Hullebusch, E.D. & Mineral resources 399–400
Annachhatre, A.P. 180 Mining financing options 419–422
Jeran, P.W. and Trevits, M.A 121 Mining hazards 386–388
Johnson, M. & Nilsson, K. 172 Mining method selection 198, 230–234
Joint Ore Reserves Committee 399–405 Mining risk and risk management 385–398
Mining Rock Mass Rating 46, 209–211
Karmis, M.,Godman, G & Hasenfus, G. 174 Mining Science & Technology, University of
Kempson, W.J, Webber-Youngman, R.C.W. Wollongong 114–117
& Meyer, J.P. 221 Mining valuation and financing 399–425
Kennedy, C. 75–79 Missavage, R. 81–89
Key block 144–145, 308–309 MIT Mission 346–347
Kim, J.S., Lee, K.S., Cho, W.J., Choi, H.J. & Mkandawire, M. & Oakes, K. 361
Cho, G.C. 181 Mohr-Coulomb equation 42
Kirkey, J. 347 Molinda, G., Mark, C., Bauer, E.R., Babich,
Kirsch, E.G. 85–88 D.R. & Pappas, D.M. 156
Kostescki, T.R. & Spearing, A.J.S. 134 Monsalve, J.J., Soni, A., Ripepi, N. &
Kratzsch, H. 131 Rodriguez-Marek, A. 99
Krige, D.G. 216 Morelli, G.H. 202
Kuchling, K.J. 357 Mountaintop mining 21–23
Multi-cut coal mining 142–144
Lacy, P. & Rutqvist, J. 353–354
Lacy, W. 187–180 Naganathan, S., Subramaniam, N. &
Lame, G. 87 Nasharuddin Bin Mustapha, K. 180
Laubscher, D.H. 210–211, 250 National Coal Board 174
Lee, F.T. & Abel, J.F. 129–130 National Institute of Occupational Safety
Lignite 60 and Health 163
Lilford, E.V. 404, 408–410 Nicholson, R.V., Ludgate, I., Clyde, E. &
Lilford, E.V., Maybee, B. & Packey, D. 418 Venhuis, M. 376
Longwall hard rock tabular mining 235–241
Longwall mining 101–130 Oberle, B. 374–375
Longwall plough(plow) and shearer 107–110 Open stoping 256–262
Longwall shields 110–114 Openpit mine design process 39–40
Lynch, S. 164 Openpit mining 28–39
Openpit planning and operation 35–38
Main access development 226–229 Ortlepp, W.D.O. 305
Mark, C. 146–147 Ostrowski, W.J.S. 83–84
Mark, C. & Agioutantis, Z. 101 Ostensson, O. & Roe, A. 360–361
Mark, C. & Gauna, M. 180–181 Overall wall slope and angle design 32
Mark, C. & Molinda, G.M. 67, 70–71 Owen, J.R., Kemp, D., Lebre, E., Svobodova,
Market risk 415–417 K. & Perez Murillo, G. 373
Martin, T.E. & McRoberts, E. 368–369 Ozbay, M.U. & Roberts, M.K.C. 93
Mathematical design 46–51
Matthews, E, & Spearing, A.J.S. 366 Palarski, J. 142–144
McDowall, W., Geng, Y., Huang, B., Panek, L. 144–145
Bartekova, E., Bleischwitz, R., Turkeli, S., Parizek, R.R., & Ramani, R.V. 176
Kemp, R. & Domenech, T. 352–353 Peat 59–60
McMullen, L. 389–392 Peck, R.B. 88–89
Medkova, K. 378 Peng, S.S. 89–90, 101–104, 117–120, 122–131
Meilani 384 Permitted explosives 287
Melbye, T. 329 Pillar factors of safety 98–100
Michaud, D. 99–100, 278 Pitwall stability monitoring 38
Mine access closure 166–167 Plate theory 152–153
Mine Safety & Health Administration 299 Pongpanya, P., Sasaoka, T., Shimada,
Mine Subsidence Engineering Consultants 128 H., Hamanka, A. & Wahyud, S. 89–90
Mine waste disposal 364–376 Prajapati, G. 34
Index 431

Pressure arch 310 Space mining 397–398


Preston, C. 294 Spearing, A.J.S. 142, 159, 235–241, 333, 335
Price Waterhouse Cooper 351–352 Spearing, A.J.S., Benton, D., Kostecki,
Project risk 418–419 T.R. & Hirschi, J.C. 133
Proust, A. 140 Spearing, A.J.S., Tadolini, S.C. & Kostecki,
Pugh, G.M. & Rasmussen, D.G. 273 T.R. 146–156
Speers, C.R. & Spearing, A.J.S. 155
Q System 200–206 Specialized coal mining methods 130–144
Quarrying 15–20 Spiling bolts 323
Queens University 245, 270 Stephenson, P. 405
Stillborg, B. 205–206, 308–309, 311, 314–317
Raffaldi, M.J., Seymour, J.B., Abraham, Strip mining 13–15, 20–21
H., Zahl, E. & Board, M. 274 Sub-level caving 242–246
Raises 222–225 Sub-level stoping 256–262
Raise Boring 223 Support design and application 144–161
Rauch, H.W. 177 Support design methodologies 146–155
Reclamation and rehabilitation 27 Sun, H., Ma, L., Lui, W., Spearing, A.J.S.,
Rehabilitation 38–39 Han, J. & Fu, Y. 181
Reinforced beam 310 Surawar, S. 57
Rhodes, A. 385 Suspension 144–145
Rio Tinto 395–397 Sustainable Development Goals 343–345
Risk definition 386 Sustainable development of minerals 357–364
Risk management in coal mines 162–163
Roadway design 34–35 Tailings dam future operations 374–376
Robinsky, E.I. 371–372 Tailings major hazards 373
Rock bolts and cable bolts 313–320 Tannant, D. & Regensburg, B. 35
Rock Mass Rating 46, 206–209 Taylor, D. 49–51
Rock Quality Designation 46, 199–200 Terminology 4–13
Rock sampling 196 Terzaghi, K. 46–47, 82–84
Rodriquez, R. 195 The Joint Ore Reserves Committee 399–405
Room and pillar mine design 91–101 Time value of money 413–415
Roseke, B. 78–79 Timoshenko, S. & Woinowsky-Krieger, S.
Row by row blasting 25–26 151–153
Rowland, J.H. & Mainero, R. 291–292 Tiseo, I. 378–379
Top coaling 139–141
Salamon, M.D.G. 230 Topf, A. 346
Salamon, M.D.G. & Munro, A.H. 93–94,97 Tributary area theory 95–97
Sawing 18 Trongvann, N. 53–54
Schissler, A.P. 23–24 Tucker, H., Holder, A., Swarts, B., van Strijp,
Schwab, K. 351 T. & Grobler, E. 304
Seedsman, R. 160 Types of rock drills 294–296
Self-drilling anchors 322 Tyson, R. 362–363
Shafts and declines 218–220
Shotcrete 323–333 Ulmer-Scholle, D.S. 346
Shi, G.H. 308 Ultra-high seams with single cut longwalling
Shrinkage mining 274–278 141–142
Singh, K.B & Dhar, B.B. 173–174 Undercut and fill mining 271–274
Singh, M. 175 University of Wollongong 314
Slope design 44 Unsupported mining methods 233–251
Slope monitoring 53–54 Urban mining 377–384
Social license to operate 362–364 US Corps of Engineers 153–154
Society of Mining Professors 397–398 US Environmental Protection Agency 179–180
Soft rock mining 56–181
Solid Backfill Mining 135–139 Vandewalle, M. 327
Soliman, M.M. 173 Vardan, H. & Murthy, S.N. 295
Solymos, A. 140 Vertical Crater Retreat 278–281
Somagani, P.C. 4,33 Vick, S. 367–368
Sovereign risk 417–418 Vietti, A. 375–376
432 Index

Wagner, D. 94–95, 97 Wire cutting 15–17


Wagner, H. 230 Wiseman, N. 312
Warner, A. W. 230–233 World Economic Forum 346, 348
Warner, R.K. 1–3 World Mining Data 350
Waste dam operation and monitoring 370–373 Worsey, P.N. 302
Waste dam rehabilitation 376
Water conservation and subsidence control Yielding rockbolts 320–322
172–179
Wedge cut 301–302 Zhang, J., Li, M., Taheri, A., Zhang, W.,
Westcott, P., Pitkin, G. & Aspinall, T. 5 Wu, Z. & Song, W. 135,168
Whittaker, B.N. & Reddish, D.J. 132–133, Zhang, Q. & Wang, X. 225–226, 229, 244,
174–175 253, 257, 259, 281
Williams, T.J., Brady, T.M., Bayer, D.C, Zhang, Z.X. 286
Bren, M.J., Paklanis, R.T., Majersison, Zhao, Y.X. & Jiang, Y.D. 181
R.T. & Langston, R.B. 272 Zhou, J., Shi, X. & Li, X. 305
Willie, D.C. & Mahn, C.W. 53 Zhou, Y., Zhang, D., Fan, G. &
Wilson, J.W. & More O’Ferrall, R.C. 230 Zhang, S. 140–141

You might also like