You are on page 1of 284

Rockburst Support

2018 ©

Reference Book

Volume 1
Rockburst Support – Volume 1

Rockburst Phenomenon and Support Characteristics

Ming Cai and Peter K Kaiser

2018 ©

Hold
Yield

Reinforce Reinforce
Retain
Cai and Kaiser

Hold
Rockburst Support Reference Book (I) 3

Rockburst Support
Reference Book
Volume I: Rockburst phenomenon and support
characteristics

By Ming Cai and Peter K Kaiser


MIRARCO – Mining Innovation, Laurentian University, Sudbury,
Ontario, Canada

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

© 2018

Draft manuscript – Copyright protected – Cai and Kaiser 2018


4 Introduction
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

MIRARCO at Laurentian University is publishing the Rockburst Support


Reference Book as a major update of the 1996 Canadian Rockburst
Support Handbook (Kaiser et al. 1996).
Three volumes will be released in 2018:
I. Rockburst phenomenon and support characteristics
II. Rock support to mitigate rockburst damage caused by dynamic
excavation failures
III. Rock support to mitigate rockburst damage caused or dominated
by dynamic disturbances from remote seismicity
Preliminary manuscripts will be distributed to interested parties offering
timely feedback and contributions for possible inclusion. Final copies
will be distributed free upon registration as a reader.

ISBN 978-0-88667-096-2 (paperback)


ISBN 978-0-88667-099-3 (Three-Volume Set paperback)
Rockburst Support Reference Book – Volume I: Rockburst phenomenon
and support characteristics
Copyright © 2018 by the authors. All rights reserved.
This book is distributed free of charge upon individual registration with
the expectation that it will find wide application but without any warran-
ty.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 5

Preamble
The Canadian Rockburst Program was completed in 1995 and the
Canadian Rockburst Support Handbook (CRBSHB) was published
in 1996 (Kaiser et al. 1996). The Canadian Rockburst Support
Handbook presented an engineering approach for the selection
rock support systems for burst-prone mines by systematically
assessing both support demands and support capacities.
This book evolved from the Canadian Rockburst Support Hand-
book and includes findings from subsequent research by the
authors and others around the world. It also summarizes experi-
ences gained over many decades and builds on the lecture present-
ed by Kaiser and Cai (2013) at the 7th International Symposium on
Ground Support in Mining and Underground Construction in Perth
(Australia) entitled ”Critical review of design principles for rock
support in burst-prone ground – Time to rethink!” In an effort to

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
overcome several deficiencies identified in 2013, this book specif-
ically addresses support design aspects for strainbursts and for
conditions where it is difficult, if not impossible, to establish a
priori the kinetic energy demand. Deformation-based support
design principles are introduced to assist in selecting robust
support systems to mitigate rockburst damage.
In recent years, many new rock support components have been
developed to enhance the ground control toolbox and new labora-
tory and field test data have become available to better assess the
dynamic performance of rock support. Data processing tools were
developed to better assess the risk of rockburst and rockburst
damage. Furthermore, it was learned that one of the most im-
portant rockburst damage processes, called dynamically triggered
strainburst, often dominates the failure process of excavations
when statically loaded by mining-induced stresses and dynamical-
ly triggered by a remote seismic event – a process that was not
fully reflected in the engineering approach presented in Kaiser et
al. (1996).
After more than a decade, it was deemed necessary to update the
handbook in the form of a reference book for rockburst resistant
support selection to assist practicing engineers in understanding
excavation damage processes and in following systematic support
design procedures. A new design aspect, the concept of seismical-
ly triggered and dynamically loaded strainbursts, is introduced.
Even though it is realized that further research will be required to

Draft manuscript – Copyright protected – Cai and Kaiser 2018


6 Introduction

fully verify the proposed design procedures, this important aspect


is now included in a qualitative and semi-quantitative manner in
this revision. This book provides comprehensive guidance to
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

create safe working conditions in burst-prone mines.


Due to the complex behaviour of highly and dynamically stressed
rock in hard rock mines, it is difficult to provide strict guidelines.
Hence, this reference book is not a “cook book” and the reader is
expected to respect site-specific experiences when following the
outlined approach. Nevertheless, this reference book is a source of
information that should assist in making prudent decisions.

This Rock Support Reference Book consists of three volumes:


- Volume I - Rockburst phenomenon and support characteristics
- Volume II - Rock support to mitigate rockburst damage caused by dynamic
excavation failures
- Volume III - Rock support to mitigate rockburst damage caused or domi-
nated by dynamic disturbances from remote seismicity

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 7

Acknowledgements
First of all, the authors wish to recognize the contributions of the
co-authors D. Tannant and D. McCreath of the Canadian Rock-
burst Support Handbook as well as the contributions of the spon-
sors of the Canadian Rockburst Program led by CAMIRO be-
tween 1990 and 1995. This work laid the foundation for this
reference book and contributed directly and indirectly to many
sections of this guide. Relevant sections of the CRBSHB’96 are
reproduced, with or without modification where deemed appropri-
ate. If we have used wordings from the original version, it is
because we could not write it better even a decade later.
The authors also wish to acknowledge the financial and substantial
technical in-kind contributions from various research sponsors
over the last decade, namely Freeport-McMoRan, LKAB, New-
crest, Rio Tinto, Vale, CEMI, and MIRARCO. The authors wish

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
to thank D. Duff for managing the progress of the project and
coordinating with industry sponsors. We also acknowledge the
substantial financial contributions from the National Sciences and
Engineering Research Council (NSERC) of Canada and the
Ontario Research Fund (ORF) in support of research that contrib-
uted to the technical content of this reference book.
The authors are grateful to D. Thibodeau, L. Malmgren, B.
Woldemedhin, A. Punkkinen, S. Maloney, S. Nickson, M. Yao, B.
Valley, N. Bahrani, N. Golchinfar, C. Groccia, X. Wang, and A.
Manouchehrian, amongst others for their contribution to this
project. In particular, we thank D. Thibodeau, L. Malmgren, B.
Woldemedhin for taking time to read an early research report and
for providing feedback and suggestions which greatly improved
the content of this reference book.
Finally, the authors wish to thank D. McCreath, T.R. Stacey, M.
Diederichs, G. Russo for taking time to critically review the early
manuscript and provide constructive criticism and suggestions that
greatly improved the quality of this work. The authors also want to
thank Kathi Kaiser for proof reading the manuscript.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Draft manuscript – Copyright protected – Cai and Kaiser 2018
8 Introduction
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Rockburst Support Reference Book (I) 9

Table of Content
Preamble ········································································· 5
Acknowledgements ···························································· 7
Table of content ································································ 9
1 Introduction to Rockburst Support Reference Book ············· 11
1.1 Background ································································· 11
1.2 Need to overhaul the rockburst support handbook? What has
changed since the 1990s? ····················································· 12
1.3 Scope of rockburst support reference book ···························· 19
1.4 Layout of reference book ················································· 20

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
1.5 References ·································································· 24
2 Rockburst phenomenon and rockburst damage ···················· 28
2.1 Rockburst phenomenon ··················································· 28
2.2 Types of rockbursts························································ 39
2.3 Mine seismicity causing dynamic disturbances ······················· 53
2.4 Rockburst damage mechanisms ········································· 58
2.5 Factors influencing rockburst damage·································· 73
2.6 Rockburst damage severity··············································· 77
2.7 References ·································································· 85
3 Design principles and methodology ···································· 93
3.1 Engineering principles ···················································· 92
3.2 Support and its function in stress-fractured ground ·················· 96
3.3 Rockburst support design principles ···································· 103
3.4 Support design methodology············································· 109
3.5 Overview of rockburst support design process························ 113
3.6 Mitigation of rockburst damage caused by excavation failure and
dynamic disturbances ·························································· 116
3.7 References ·································································· 119

Draft manuscript – Copyright protected – Cai and Kaiser 2018


10 Introduction

4 Capacity of rock support components ································· 123


4.1 Characteristics of rock support elements······························· 123
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

4.2 Rock support element testing – pull and drop tests··················· 129
4.3 Rockbolt test results ······················································· 143
4.4 Test results of surface support components ··························· 146
4.5 Summary of rock support component capacities ····················· 152
4.6 Suggested design capacities for support design ······················· 157
4.7 References ·································································· 167
5 Rock support system capacity ··········································· 173
5.1 Rockburst damage mitigation············································ 173
5.2 Integrated support system characteristics ······························ 175
5.3 Estimation of support system capacity ································· 186
5.4 Rock support systems – ground-truthing······························· 215
5.5 References ·································································· 232
Appendix A: Terminology ···················································· 235
Appendix B: Nomenclature··················································· 241
Appendix C: Static and dynamic capacities of rockbolts ················ 245
Appendix D: Static and dynamic capacities of surface support compo-
nents ·············································································· 249
Appendix E: Information sheets for rock support components ········· 254

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 11

Chapter One
Introduction to Rockburst
Support Reference Book

1 Introduction to Rockburst Support Reference Book ................... 11


1.1 Background ......................................................................... 11
1.2 Need to overhaul the rockburst support handbook? What has
changed since the 1990s? ................................................................ 12

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
1.2.1 Background .................................................................. 12
1.2.2 Recent qualitative observations from Reservas Norte.... 18
1.3 Scope of rockburst support reference book ........................... 19
1.4 Layout of reference book ..................................................... 20
1.5 References ........................................................................... 24

1 Introduction to
Rockburst Support Reference Book
1.1 Background
In the 1990’s, as part of the Canadian Rockburst Research Pro-
gram, the Geomechanics Research Centre (GRC) at Laurentian
University, Sudbury, Ontario, Canada, undertook an extensive
research program aimed at providing a rational design methodolo-
gy for selecting support systems in burst-prone ground. This work
was funded by the Mining Research Directorate (MRD), repre-
senting major Canadian and foreign mining companies, and by the
Ontario Ministry of Northern Development and Mines. Funda-
mental research associated with key issues was undertaken by
graduate students funded by the Natural Sciences and Engineering
Research Council (NSERC) of Canada. The key conclusions of
this research were summarized in the Canadian Rockburst Support
Handbook (Kaiser et al. 1996) and the purpose of this handbook

Draft manuscript – Copyright protected – Cai and Kaiser 2018


12 Introduction

was to assist practitioners in selecting support systems to control


and limit damage caused by rockbursts.
The behavioural complexity of rock mass and tunnel support
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

systems under dynamic loading conditions does not permit to


provide a simple, prescriptive cookbook. Rather, it is necessary to
provide an orderly approach to the problem, to simplify and bring
into focus the essential mechanisms involved before providing
specific guidance on the selection of appropriate support systems.
Because many simplifications and assumptions had and still have
to be made, the authors emphasize the need for careful judgment
when using this reference book and the related engineering tools.
This reference book is intended for ground control engineers with
a sound background in rock mechanics, stress modelling, engi-
neering design, and static ground control practices. Thus, a basic
understanding of rock mechanics and a sound understanding of
some ground control tools such as bolts, mesh, and shotcrete are a
pre-requisite. This reference book will be of greatest value when
combined with practical experience from mining in highly stressed
ground.

1.2 Need to overhaul the rockburst support hand-


book? What has changed since the 1990s?
1.2.1 Background
Codelco was a supporting member of the Canadian Rockburst
Research Program (1990-95) and the experience from the mining
block TEN-SUB-6 at the El Teniente Mine in Chile led to a
support design rationale that was heavily influenced by severe
energy releases from large seismic events ranging from ML = 3.2
to 4.0 (Richter magnitude). As a consequence, the design rationale
published in the Canadian Rockburst Support Handbook
(CRBSHB; Kaiser et al. 1996) was ‘ground-motion-centric’ in that
the rockburst damage was primarily related to the energy radiated
from remote, mostly fault-slip type, seismic sources. It was as-
sumed that the stress waves and the associated peak ground
velocities (PGV) or accelerations (PGA) or dynamic stress pulses
(Dsd ) were the primary causes for the severity of rockburst dam-
age (i.e., for the observed ejection intensity).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 13

The TEN-SUB-6 mining block experienced numerous major


rockbursts between 1990 and 1992 when it had to be shut down
due to frequent and severe rockbursting (Figure 1-1; red arrow).
Since then, the mine has made major changes to its operating
procedures, introduced hydro-fracturing as means to reduce the
intensity of seismicity, and improved the ground support by
adding substantial energy dissipation capacity to the support
systems. As a consequence, the frequency of seismic events and
the associated rockbursts were reduced (Figure 1-1).

Hydrofracturing

# rockbursts

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Figure 1-1 Rockburst history at El Teniente Mine from 1992 to 2015 (Joo 2017).
KTPD – 1000 tons per day (production rate).

The following excerpts from Chapter 9 of the CRBSHB (Kaiser et


al. 1996) are reproduced here to highlight aspects that dominated
the research and engineering thinking at the time.

El Teniente SUB-6
The TEN-SUB-6 production area is located to the northeast of the
Braden Pipe, a chimney of subvolcanic breccia that post-dated the
principal copper/molybdenum mineralization (Tinucci and Trifu
1994). At SUB-6, the porphyry copper orebody consists primarily
of andesite with diorite and porphyritic andesite intrusions. Prima-
ry ore is extracted using the block caving method and mining of
the upper TTE-1 and TTE-4 levels had resulted in a large caved
area above SUB-6 (Figure 1-2).
Severe rockburst problems plagued the El Teniente SUB-6 pro-
duction level (2105 m) since production began in August 1989.
Some large seismic events (Table 1-1 and Figure 1-3) resulted in

Draft manuscript – Copyright protected – Cai and Kaiser 2018


14 Introduction

excavation and support damage ranging from minor shotcrete


spalling to severe support and rockmass collapse (and, in most
cases, floor heave) requiring extensive drift rehabilitation.
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

3600-3800 m

Slope Events

TTE-1 (2624 m)
Caved Area

Toe Events
Braden TTE-4 (2354 m)
Pipe
Caving Events

SUB-6 (2105 m)

Production Events

Deep Events

Figure 1-2 Location of areas with strong fault-slip like seismic energy release
(Kaiser et al. 1996; Chapter 9).

Table 1-1 Major seismic events recorded at El Teniente SUB-6 from January
1990 to June 1992 (Table 9.7; CRBSHB 1996)

Event Date Richter magnitude ML*


1 18 January 1990 3.6
2 2 July 1990 3.2
3 19 April 1991 -
4 24 April 1991 -
5 23 May 1991 4.0
6 17 August 1991 -
7 17 November 1991 3.8
8 31 December 1991 3.4
9 5 February 1992 2.7
10 8 March 1992 3.6
11 25 March 1992 3.7

Note: * Estimated by the Department of Geophysics, University of Chile,


Santiago

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 15

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Legend:
1. January 18, 1990 at pickhammer level of TEN-7
2. July 2, 1990 at production level
3. April 19, 1991 at production level
4. April 24, 1991 at production and undercut levels
5. May 23, 1991 at production, caving, pickhammer, and injection levels
6. November 17, 1991 at production, pickhammer levels and TEN-7
7. December 31, 1991 at production and caving levels
8. March 8, 1992 at production and pickhammer levels
9. March 25, 2018 at production, ventilation, and pickhammer levels.

Figure 1-3 Areas of damage recorded on production level (Kaiser et al. 1996;
Figure 9.2).

Production at SUB-6 was halted following the 25 March 1992


seismic event which caused extensive damage to the production,
ventilation, and pickhammer levels. Low volume production at the
western end of the undercut level was resumed in January 1994 (R.
Dunlop, pers. comm. 1995).
Due to the exceptional rockburst damage conditions during the
period from January 1990 to June 1992, El Teniente SUB-6 was
chosen as a case study for the rockburst support part of the Cana-
dian Rockburst Research Program. Part of this rockburst investi-
gation involved modelling of the dynamic damage processes
around horseshoe-shaped tunnels excavated at SUB-6 (Vasak and
Draft manuscript – Copyright protected – Cai and Kaiser 2018
16 Introduction

Kaiser 1995a, b). The concepts of rock baggage and depth of


fracture (Chapter 6 in Kaiser et al. (1996)), identified by numerical
modelling and verified by field observations, were simplified and
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

incorporated into the overall design philosophy and into a Mine


Map Overlay (MMO) approach. Chapter 9 of the CRBSHB
illustrates the MMO concept on mining block SUB-6 at El Ten-
iente Mine.
Mine Map Overlay concept
The SUB-6 MMO model was developed to test and verify the
design concepts introduced in the CRBSHB and important aspects
required to prepare the MMOs for support design were described
in Chapter 9. The 1996 version covered the influence of mine
layout, tunnel geometry, geology, and local pre-mining stress
conditions. Results could be visualized in plans as illustrated by
Figure 1-4.

Figure 1-4 Dynamic strength factor for tunnels supported with grouted rebar
(Figure 9.13 in Kaiser et al. 1996).

As indicated above, the MMO approach was ‘ground-motion


centric’ in that it was assumed that the ground motions (PGV)
were representative of the damaging energy. At the time, the
strainburst potential (Figure 1-5) was considered as a measure of
how close the excavation was to failure. This map, although
simplistic, indicates areas with low strength factor and thus pre-
sents one of many factors that affect the strainburst vulnerability.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 17

Figure 1-5 Strainburst strength factor for the production level (Figure 9.8 in

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Kaiser et al. 1996)

This approach allowed to establish areas of elevated strainburst


potential. However, the contribution of stored strain energy or the
influence of variations in the deformation potential (mine stiffness)
was not considered in 1996 when assessing the rockburst severity.
It was assumed that dynamic stress waves could deepen the depth
of fracture and that the energy release causing excavation damage
was only related to the energy emitted from the damage causing
remote seismic source.
Furthermore, in the 1996 design approach, it was not recognized
that some of the support capacities could already have been
consumed at the time when a damaging seismic event loaded or
strained the surrounding ground. In other words, the concept of
support capacity consumption resulting from mining-induced
displacements and from repeated co-seismic deformations (seis-
mic ‘hammering’) was not considered at the time. Hence, three
key elements are missing in the CRBSHB and additional mine
map overlays have to be considered for support design purposes:
- Mining-induced displacement maps (convergence maps);
- Cumulative co-seismic support displacement maps; and
- Mine stiffness or deformation potential maps.
These aspects are introduced in Volume II (Kaiser and Cai 2018)
of this reference book.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


18 Introduction

Finally, whereas it was pointed out that a support system is only as


strong as its weakest link, means to establish the capacity of the
installed support system, consisting of various support compo-
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

nents, were not provided. Furthermore, means to establish the


remnant support system capacity, i.e., the capacity that is actually
available after the support has been deformed during previous
mining activities, were not provided.
These aspects are also covered in this reference book (Volumes I
and II).

1.2.2 Recent qualitative observations from Reservas Norte


During an underground visit of El Teniente Mine on November 18,
2017, after RASIM9-2017, it was possible to visit the mining
block Reservas Norte that contains the former TEN-SUB-6 area at
its eastern edge.
The ‘production events’ (Figure 1-2) were at the time (in the
1990’s) recognized as the cause for damage because the excava-
tions were strainburst-prone but, as indicated above, the impact of
the stored strain energy release on the support design was ignored
for the development of the CRBSHB. As a consequence, the
support design approach was dominated by the assumption that
energy from ‘remote’ seismic events is the primary or sole source
of damage causing energy. Today, it is recognized that a sound
support design also has to consider the often-dominating energy
that is released from the rock mass surrounding the ‘burst vol-
ume’1. Even though this does not seem to be formally recognised
in the currently adopted support design approach for El Teniente
Mine, the steady push for more energy capacity in the support
system suggests that the mine has recognized that ‘ground-motion
centric’ design alone is not sufficient to ensure excavation stability
in areas with high rockburst hazards. Their heavy bumper support
design also highlights that dynamic support at the lower corners of
tunnels, between wall and floor, is essential to maintain the overall
excavation stability in severely bursting ground.
Finally, even though discussions on site with the mine engineers
suggest that El Teniente Mine has not formally adopted a preven-

1
The ‘burst volume’ is the volume of rock that suddenly fails (bursts) and rapidly moves
toward the excavation.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 19

tive support maintenance practice, it is quite evident that they are


adopting a proactive support enhancement approach. For example:
- in the Dacita formation, known to be very brittle and thus highly
strainburst-prone, a two-layer support system with high strength
chain-link (Geobrugg) mesh is used throughout; and
- in the extraction zone (i.e., in the high stress transition zone), the
second layer of the two-layer support system with high strength
chain-link mesh is only installed in areas of anticipated excessive
displacement and energy release; e.g., in fault zones that experienced
rockbursts in the past and in areas with heavily cracked or debonded
shotcrete.
These observations confirm that major adjustments to the
CRBSHB are needed to account for the impact of the strainburst
energy. Most importantly, the observed support damage confirms
that excessive displacement due to rock mass bulking can cause

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
damage to support that is otherwise effective in preventing rock
ejection. This suggests a shift toward a deformation-based (rather
than energy-based) support design approach (introduced in Vol-
ume II) and a proactive support maintenance strategy to manage
the remnant support system capacity. These aspects are now fully
covered in this reference book.

1.3 Scope of rockburst support reference book


This book offers a state-of-the-art methodology for support selec-
tion to ensure work place safety. Definitions are presented in the
‘Terminology’ section (Appendix A, Volume I) to prevent confu-
sion that may arise from alternate uses in the literature. For exam-
ple, a rockburst is defined as “damage to an excavation that occurs
in a sudden or violent manner and is associated with a mining-
induced seismic event.” Even though the term rockburst implies
damage, we refer to ‘rockburst damage’ when specifically discuss-
ing the damage process, its extent and severity. The emphasis in
this book is on the damage caused by the rock mass failure near an
excavation rather than on a seismic event that may or may not be
the dominant energy source for the damage. This book then
provides a methodology for rockburst damage assessment and risk
mitigation by ground support.
In this reference book, the primary focus is on rock mass behav-
iour at the damage location itself because it is necessary to under-
stand the mechanisms that actually cause damage. Of course, it is
Draft manuscript – Copyright protected – Cai and Kaiser 2018
20 Introduction

also necessary to understand how those mechanisms may have


been triggered or affected by dynamic loading. The specific
damage mechanism and the severity of the damage will vary
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

depending upon a number of factors such as the pre-existing


mining-induced stress level, the local geology and quality of the
rock mass, the excavation shape, the seismic source characteristics,
and many more. These two broad aspects, i.e., the nature of the
damage mechanism and the severity of the resulting damage,
when combined, define the demand which will be placed upon an
installed rock support system and the reinforced rock mass.
There are three types of demand that will be considered: (1) the
load demand – how much weight the support needs to carry; (2)
the displacement demand – how much the support must be able to
give (deform) without failing; and (3) the energy demand – how
much energy the support system and the supported rock mass must
be able to dissipate or absorb. Once the nature of these demands is
understood, it becomes possible to select a support system in a
rational manner.

1.4 Layout of reference book


Overall, this book follows the methodology introduced in the 1996
handbook with one major addition (Volume II) and several signif-
icant improvements. First, the section on support capacity has
been updated to include the characteristics of tools in a much-
improved support toolbox with many new support system compo-
nents (Volume I). Second, the reference book now differentiates
between seismically triggered and dynamically loaded strainbursts
and offers a deformation-based support selection approach (Vol-
ume II). Third, for conditions with very strong ground motions,
the third volume (III) summarizes those aspects that are still valid
for support selection. For this purpose, two software tools were
developed to assist in in the application of the ‘ground motion-
centric’ approach. Guidance for forensic analyses of rockburst
damage is provided in Volume III as a means to properly identify
and assess the sources of damage and the underlying damage mecha-
nisms.

Figure 1-6 illustrates the layout and content of the reference book
with three volumes and twelve chapters.
- Volume I: Rockburst phenomenon and support characteristics.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 21

- Volume II: Rock support to mitigate rockburst damage caused by


dynamic excavation failures.
- Volume III: Rock support to mitigate rockburst damage caused or
dominated by dynamic disturbances from remote seismicity.

Volume I
Chapter 1 – Introduction
Chapter 2 – Rockburst phenomenon and rockburst damage
Chapter 3 – Design principles and methodology
Chapter 4 – Capacity of support components
Chapter 5 – Rock support system capacity
Appendix A – Terminology
Appendix B – Nomenclature
Appendix C – Static and dynamic capacities of rockbolts
Appendix D – Static and dynamic capacities of surface

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
support components
Appendix E – Rock support element info sheet

Volume II
Chapter 1 – Excavation behaviour and vulnerability to
rockburst damage
Chapter 2 – Strainbursts
Chapter 3 – Deformation-based support design

Volume III
Chapter 1 – Assessment of seismic risk and hazard for
ground control planning
Chapter 2 – Seismic shakedown – acceleration based design
Chapter 3 – Rock ejection – ground motion based design
Chapter 4 – Support system selection
Chapter 5 – Design verification and modification

Figure 1-6 Content of Rockburst Support Reference Book.

In Volume I, the rockburst phenomenon and the damage processes


with measures of damage severity are introduced in Chapter 2.
Rockbursts are often classified in terms of the event magnitude or
seismic moment but for a support design, the extent of damage in
terms of depth of failure and lateral extent matters. Hence, the
severity of damage needs to be defined for support design purpos-
Draft manuscript – Copyright protected – Cai and Kaiser 2018
22 Introduction

es in terms of rockburst damage characteristics (induced forces


and deformations). Second, engineering principles and approaches
are presented in Chapter 3 with an overview of the structure of the
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

support selection process. A sound comprehension of the underly-


ing methodology is important to guide the ground control engi-
neers through the selection process of support systems for burst-
prone ground. For this, it is necessary to know what types of rock
support technologies are available, what the capacity of each
support element is, and how to combine them into an integrated
support system. This is summarised in Chapter 4 and a detailed
summary of currently available rock support technologies is
presented in Appendix C of Volume I.
A detailed discussion of the excavation vulnerability is presented
in Volume II (Chapter 1) because vulnerable excavations are more
susceptible to rockburst damage and prone to severe damage when
stored strain energy is released. Many observed rockbursts in
underground mines are in fact strainbursts. Some of them are self-
initiated, due to gradual mining-induced stress changes, and some
are triggered by dynamic stress waves emitted from remote seis-
mic events. For self-initiated or triggered strainbursts, the primary
or secondary seismic event is co-located with the excavation
damage, i.e., the damage is the source of energy release. Damage
to the excavation is associated with a seismic event but is not
dominated by the stress wave emitted from the seismic source.
Excavation damage is the source mechanism. Volume II is thus
devoted to the mitigation of rockburst damage caused by dynamic
excavation failures. In particular, a detailed discussion of strain-
bursts and the associated deformation and energy release process-
es is presented here. Because it is difficult, if not impossible, to
estimate the kinetic energy released by strainbursts, a defor-
mation-based support design methodology is introduced to man-
age stress-fractured ground resulting from static and dynamic rock
mass failure. Means for the assessment and quantification of
support capacity consumption and, as a consequence, proactive
support maintenance are also described in Volume II.
Because rockburst damage can also be caused by seismic events
that are located at some distance from an excavation, i.e., remote
from the damage location, the seismic hazard is frequently defined
in terms of ground motions emitted from a seismic event. This is
discussed in Volume III to form the basis for dynamic stress and
acceleration determination. Volume III also describes techniques
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 23

by which detailed analyses of the ground motion patterns can be


undertaken. Aspects of ground motion assessment are now sup-
ported by a synthetic ground motion assessment tool (S-GMAT
software; CEMI/IMS, 2012). The reference book then completes
the support design methodology by focussing on failure processes
dominated by support demands related to remote seismic events.
Specifically, Volume III covers support design aspects to prevent
seismic shakedown resulting from dynamic disturbances causing
incremental accelerations and covers aspects of excavation dam-
age dominated by ground motions from large remote seismic
events.
Practical implications and applications as well as case histories are
discussed throughout the book to illustrate the applicability of the
proposed approaches and to guide the user in selecting rock
support for burst-prone ground.

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Supporting software and spreadsheets
Several elements of support design described in this book are
supported by a software package called BurstSupport (Cai et al.
2012). This tool simplifies the use of the reference book for those
aspects covered in Volume III and augments the ground control
engineer's ability for making day-to-day operational decisions
regarding excavation damage potential and support performance
in situations dominated by large seismic events. It allows practi-
tioners to apply the charts and figures contained in the reference
book and to adjust them for site-specific conditions. This tool does
not yet cover support design aspects of strainburst-dominated
failure processes but supporting spreadsheets are offered to assist
with the application of the approaches presented in Volume II. It is
hoped that funding will become available in the future to expand
the mine map overlay approach to account for strainburst damage
and associated support design principles.
Finally, whereas many sections of the CRBSHB have now been
revised and supplemented with advanced support selection princi-
ples, the reader is encouraged to consult the CRBSHB (Kaiser et
al. 1996) because it contains valuable case studies and test results
that are not repeated in this reference book.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


24 Introduction

1.5 References
Cai, M., Kaiser, P.K., and Duff, D. 2012. Rock support design in burst-prone
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

ground utilizing an interactive design tool. In 46th US Symp. Rock Mech,


Chicago. Paper 12-599 p. 8.
Joo, A.M. 2017. Dynamic ground support at El Teniente Mine. Personal
communication with P.K. Kaiser. RASIM9 short course presentation at
RASIM9.
Kaiser, P.K., and Cai, M. 2013. Critical review of design principles for rock
support in burst-prone ground - time to rethink! Keynote. In Ground
Support 2013. Edited by Y. Potvin and B. Brady. pp. 3-38.
Kaiser, P.K., and Cai, M. 2018. Rockburst support reference book. Volume
II: Rock support to mitigate rockburst damage caused by dynamic
excavation failures. MIRARCO, Laurentian University, Sudbury,
Ontario.
Kaiser, P.K., Tannant, D.D., and McCreath, D.R. 1996. Canadian Rockburst
Support Handbook. Geomechanics Research Centre, Laurentian
University, Sudbury, Ontario. p. 314.
Tinucci, J., and Trifu, C. 1994. Assessment of Mining-induced Seismic
Mechanisms at El Teniente Mine, Vol I, Task II, Report by Itasca
Consulting Group Inc. p. 80.
Vasak, P., and Kaiser, P.K. 1995a. Rock and Support Damage Assessment
for El Teniente Mine - Phase II Progress Report. Numerical Modelling
for TEN-SUB6. Codelco, El Teniente Mine, Chile.
Vasak, P., and Kaiser, P.K. 1995b. Tunnel stability assessment during
rockbursts. In CAMI'95 3rd Canadian Conference on Computer
Applications in the Mineral Industry, Montreal, Quebec. pp. 238-247.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Chapter Two
Rockburst phenomenon and
rockburst damage
Synopsis
In this reference book, a rockburst is defined as “damage to an
excavation that occurs in a sudden and violent manner and is
associated with a seismic event.” Key aspects of the rockburst
phenomenon – why rockbursts occur and how they differ with
respect to cause and effect are summarised. Three rockburst types,
i.e., fault-slip-, pillar- and strain-bursts, were discussed in the 1996
Canadian Rockburst Support Handbook (Kaiser et al. 1996). Much
more emphasis to strainbursts is given here as it was found that
excavation damage is almost always related to energy released
from the rock mass surrounding the burst volume at the damage
location. Strainbursts can be mining-induced, seismically triggered
by a remote seismic event or dynamically loaded by the related
stress pulse. Similarly, falls of ground can be triggered or dynami-
cally loaded by dynamic disturbances from remote seismic events.
For support selection purposes, it is necessary to differentiate
between four dynamic damage mechanisms: (1) sudden stress-
induced fracturing or strainbursting due to tangential straining of
the excavation; (2) rock ejection due to a high rock mass bulking
rate during strainbursting; (3) rock ejection by energy transfer
from a remote seismic source; and (4) shakedown or falls of
ground due to acceleration forces elevated by the impact of ground
motions from a remote seismic event. These mechanisms and the
factors that influence the rockburst damage potential and damage
severity are discussed. Most importantly, the relevance of an
excavation’s vulnerability to damage by seismic disturbances is
introduced in this chapter and then explained in more detail in
Volume II. Emphasis is placed on the important roles of strain-
bursting and the remnant capacity of the installed support in iden-
tifying the vulnerability to failure.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


26 Rockburst phenomena and rockburst damage

2 Rockburst phenomenon and rockburst damage ............................ 29


Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

2.1 Rockburst phenomenon ........................................................ 29


2.1.1 What is a rockburst? ..................................................... 29
2.1.2 Why rockburst occurs?.................................................. 29
2.1.3 Are there different kinds of rockbursts?......................... 30
2.1.4 Where do rockbursts occur? .......................................... 31
2.1.5 How large or severe is a rockburst? ............................... 34
2.1.6 What has been done to deal with rockburst? .................. 38
2.2 Types of rockbursts .............................................................. 40
2.2.1 Strainburst .................................................................... 41
2.2.2 Pillar bursts................................................................... 47
2.2.3 Fault-slip rockbursts ..................................................... 49
2.3 Mine seismicity causing dynamic disturbances ..................... 54
2.3.1 Seismic source characteristics................................... 55
2.3.2 Ground velocity for support design ............................... 56
2.3.3 Ground acceleration for support design ......................... 57
2.3.4 Dynamic ground stress .................................................. 58
2.4 Rockburst damage mechanisms ............................................ 59
2.4.1 Seismically induced falls of ground............................... 63
2.4.2 Rock mass bulking........................................................ 64
2.4.3 Rock mass buckling ...................................................... 65
2.4.4 Rock mass bulking enhanced by shear rupture............... 69
2.4.5 Rock ejection ................................................................ 71
2.5 Factors influencing rockburst damage ................................... 74
2.5.1 Geotechnical factors ..................................................... 74
2.5.2 Geology ........................................................................ 75
2.5.3 Mining.......................................................................... 75
2.5.4 Seismicity ..................................................................... 76
2.5.5 Combination of factors.................................................. 77
2.5.6 Dealing with rockburst damage ..................................... 77
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 27

2.6 Rockburst damage severity................................................... 78


2.6.1 Classification of damage severity ................................. 78
2.6.2 Assessment of rockburst damage severity ..................... 80
2.6.3 Ground motion tolerance of stable excavations ............. 84
2.6.4 Concluding remarks ..................................................... 86
2.7 References ........................................................................... 86

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Draft manuscript – Copyright protected – Cai and Kaiser 2018
28 Rockburst phenomena and rockburst damage
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
2 Rockburst phenomenon and rock-
burst damage

2.1 Rockburst phenomenon


2.1.1 What is a rockburst?
As the depth of mining and underground construction increases,
stress-induced failure processes are inevitable, both inside the rock
mass, away from mined openings, and near excavations. In some
cases, the rock mass fails violently leading to seismic events that
are caused by slip along weakness planes (e.g., faults) or by a
shear rupture. In other cases, the rock mass near an excavation
fractures suddenly causing strainburst damage to excavations.
Deep-seated rock mass failure or excavation failure or a combina-
tion of these mechanisms can lead to rockbursts.
In this reference book, a rockburst is defined as damage to an
excavation that occurs in a sudden and violent manner and is
associated with a seismic event (Hedley 1992; Kaiser et al. 1996).
Ortlepp (1997) gives a slightly different definition by describing
that a rockburst is a seismic event that causes violent and signifi-
cant damage to tunnels or excavations of a mine. The common
understanding is that a rockburst is associated with damage to an
excavation or its support. A seismic event alone, without causing
damage, is not a rockburst (see Appendix A Terminology).
Rockbursts can pose a substantial investment risk to operations as
they can cause damage to mine infrastructure and equipment,
injuries to workers or fatalities, resulting in rehabilitation costs
and possibly production losses. Rockburst prevention or mitiga-
tion is one of the most challenging problems in rock engineering.

2.1.2 Why rockburst occurs?


Seismic events occur when the deviatoric stress in the confined
rock mass reaches its strength and causes unstable failure by shear
slip or shear rupture. The radiated energy may then damage an
excavation and thus cause rockburst damage. Alternatively, the
deviatoric stress near an excavation may reach the unconfined or
lightly confined rock mass strength and initiate a rockburst, which
is called a strainburst with a co-located seismic event centre. Both

Draft manuscript – Copyright protected – Cai and Kaiser 2018


30 Rockburst phenomena and rockburst damage

processes cause damage to excavations and thus are rockbursts.


This reference book has been expanded to incorporate previously
ignored aspects of support selection to constrain energy releases
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

from strainbursting ground.


Many factors such as in-situ and mining-induced stresses, rock
strength, excavation geometry and sequence, and geology play a
role in triggering rockbursts. The likelihood of rockbursting or the
rockburst potential generally increases as mining progresses to
deeper ground where the in-situ and mining-induced stresses
become less favourable.
Rockbursts are mostly associated with hard brittle rock types that
are ‘intrinsically’ brittle, i.e., rocks that can store substantial
amounts of strain energy before failure and release this energy
during a rapid post-peak strength loss. If these rocks fail suddenly
in an unstable and violent manner, most of the stored strain energy
will be released. For this reason, rockbursts are frequently encoun-
tered in deep hard rock mines.
Rockbursts, however, are also associated with conditions where
energy can be stored and then released from the rock mass sur-
rounding a ‘burst volume’ (the volume of rock that fails during the
rockburst). If tractions at the boundary of a burst volume are sud-
denly removed or drastically lowered, the surrounding ground
moves in and imposes energy on the failing volume. The softer the
mine system, the more deformation follows and as a result more
energy is imposed. A low mine system stiffness renders a rock
mass as ‘relatively’ brittle referring to the post-peak strength loss
of the rock mass relative to the mine stiffness.
For this reason, rockbursts can also occur in soft rock mines such
as coal and potash mines and in situations where the mining sys-
tem is relatively soft, i.e., where the excavation has a high defor-
mation potential caused by high extraction ratios, unfavourable
geometries, the presence of weak geological structures, among
other factors. These factors are later discussed in more detail.

2.1.3 Are there different kinds of rockbursts?


In this book, as in the CRBSHB (Kaiser et al. 1996), the authors
distinguish between three fundamental rockburst types: fault-slip-,
pillar- and strain-bursts. They can be mining-induced due to
gradual static stress changes or caused by dynamic loading result-
ing from a remote seismic event or a combination of both.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 31

Fault-slip rockburst refers to damage to an excavation caused by


an energy released from a shear slip or a shear rupture source that
is remote from the excavation. Damage is caused by dynamic
disturbances emitted from the fault-slip source and may, in part or
exclusively, be related to the intensity of the seismic event which
is directly related to the source size.
A pillar burst refers to damage to an excavation caused by an
excessive loading of a pillar such that the pillar core fails violently.
Excavation damage is caused by shear rupture in the core of the
pillar. The intensity of the related seismic event and the excava-
tion damage depends on the energy released during this failure
process.
A strainburst is a sudden and violent failure of rock near an exca-
vation boundary caused by excessive straining of an un-fractured
or partially fractured volume of rock (burst volume). Because the

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
rock near an excavation is often stress fractured or plastically
deformed, strainbursting occurs near the tangential stress peak, i.e.,
it does not always occur right at the excavation wall, it can also
take place at the transition from the stress-fractured to the undam-
aged (elastic) ground. A strainburst therefore extends the depth of
stress-fractured ground or the depth of failure. The primary or a
secondary seismic source is co-located at the damage location.
The intensity of the seismic event is indicative of the burst volume,
the local stress level, and the local deformation potential.
These rockburst types are discussed in more detail in Section 2.2.
However, it is important to differentiate between when and where
rockbursts could occur, or what the rockburst potential (RBP)
might be, and how much damage they can cause, i.e., what the
rockburst severity (RBS) could be. The RBP may be low but the
RBS could be high or, vice versa, the RBP may be high but the
RBS could be low.

2.1.4 Where do rockbursts occur?


Rockbursts occur if excavations are vulnerable!
Rockbursts are violent excavation failure processes. They occur
where excavations are most prone to failure, i.e., at locations
where excavations are most vulnerable and the rockburst or
strainburst potential (RBP or SBP) is high. The vulnerability of an
excavation in terms of burst potential depends primarily on factors
that affect the proximity to failure or the strength to stress ratio,

Draft manuscript – Copyright protected – Cai and Kaiser 2018


32 Rockburst phenomena and rockburst damage

i.e., the static factor of safety (FSs; before a rockburst occurs). In


terms of rockburst or strainburst severity (RBS or SBS), the vul-
nerability depends on the amount of stored or releasable strain
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

energy, and the intensity of the ground motion emitted from a


remote seismic event. These factors will be discussed in detail in
this reference book but it can be stated that an excavation is most
vulnerable if the rock mass is close to failure, much energy is
stored in the rock mass surrounding the excavation, or is located
close to a seismic event. As a consequence, rockbursts are mostly
associated with features that increase the vulnerability to bursting,
e.g., geological structures such as dykes leading to stress raisers or
faults that affect the deformation or energy release potential.
Rockbursts occur at or near tunnel faces!
As stress redistribution, due to tunnel advance and related conver-
gence, causes incremental or gradual tangential stress changes in
the tunnel wall near the tunnel face (within two to three tunnel
diameters from the face), a rockburst can occur at or behind the
tunnel face. These rockbursts are classified as strainbursts because
stored strain energy is suddenly released when the stresses locally
reach the peak strength of the unsupported or supported rock mass.
The SBP and SBS are affected by the method of excavation and
the rate of excavation. Theoretically, the strain energy release rate
and thus the SBS should be lower during continuous excavation
with small excavation increments (Salamon 1983). However,
rockbursts may occur simultaneously with a blast and thus are not
recognized as rockbursts. For this reason, it often seems as if there
were fewer rockbursts, and thus a lesser rockburst potential, dur-
ing drill & blast tunnelling. For example, during the advance of
deep tunnels (e.g., Lötschberg and St. Gotthard tunnels in Switzer-
land, Jinping-II hydropower intake tunnels in China, and the Ol-
mos Trans-Andean tunnel in Peru) excavated by tunnel boring
machines (TBMs) minor to severe rockbursting was experienced
(Rojat et al. 2001; Robbins 2010; Zhang et al. 2012; Ma et al.
2015), whereas parallel tunnels excavated by drill & blast were
reported as less plagued by rockbursting.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 33

Figure 2-1 Rockburst damage at the Olmos Trans-Andean tunnel in Peru

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
(photo courtesy: Dick Robbins)
In tunnels excavated by a TBM, the stress changes gradually and
the rock near the excavation is minimally disturbed. Consequently,
strainbursts are affected by the advance rate and occur close to the
excavation wall. They typically strike at some distance from the
face.
Delayed strainbursts occur in situations where the stress state
remains constant but the rock strength degrades or the rock
strength is lowered over time, e.g., by a time-dependent loss of
confinement. Some delayed strainbursts at the Jinping II intake
tunnels were attributed to these processes (Feng et al. 2012).
Rockbursts behind the support or inside the reinforced rock mass
Because tangential stresses are the highest at the interface between
fractured and elastic or non-fractured rock, where the confinement
is still rather low, strainbursts can occur at some distance from the
excavation wall, i.e., inside the supported or reinforced ground. At
this interface, brittle rock is prone to stress fracturing and shear
slip along critically oriented discontinuities. The rock mass in the
burst volume shown in Figure 2-2 will therefore suddenly bulk
and impose a radial displacement d on the fractured rock and rock
support system.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


34 Rockburst phenomena and rockburst damage

Fractured rocks
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

F
Burst volume

Figure 2-2 Strainbursts are caused by the force F acting on a ‘burst volume’
located in the reinforced ground and rock mass bulking of stress-fractured
ground results in associated displacements d into the excavation.

Rockbursts cause floor heave


Rockburst often cause dynamic floor heave as the floors are rarely sup-
ported. A historic example illustrating severe dynamic floor heave in an
unsupported excavation is presented in Figure 2-3.

Figure 2-3 Dynamic floor heave in an unsupported drift (Hedley 1992).

2.1.5 How large or severe is a rockburst?


Because a rockburst is defined as damage to an excavation that
occurs in a sudden and violent manner and is associated with a
seismic event, the ‘size’ of a rockburst can be measured by the
severity of the damage to an excavation and its support, or by the
intensity of the associated energy release reflected in the seismic

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 35

event magnitude. As indicated above, in terms of rockburst or


strainburst severity (RBS or SBS), the vulnerability of an excava-
tion depends on the amount of stored or releasable strain energy,
and the intensity of the ground motion emitted from a remote
seismic event.
Rockburst damage severity
The severity of damage can be viewed from three perspectives: (1)
the volume of damaged ground (e.g., the depth of failure), (2) the
degree of damage to the installed support elements or support
system, and (3) the violence of the damage process (e.g., the seis-
mic event magnitude or the impact or ejection velocity).
(1) The rockburst damage severity is most frequently described by
the tonnage of displaced rock. It can also be characterized by
the depth and lateral extent of the fractured rock around an
opening that is involved in the failure process. For average

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
mining drift sizes, Kaiser et al. (1996) classified rockburst
damage into three severity levels: minor, moderate, and major
or severe, as shown schematically in Figure 2-4.

< 0.25 m

minor

< 1.5 m
< 0.75 m

moderate major

Figure 2-4 Rockburst damage severity (Kaiser et al. 1996).

(2) If an excavation is supported, the severity of damage to the


support system can be characterised as negligible or minor if
the support is not damaged but visually loaded, e.g., with bent
plates and loose rock in the mesh. If locally or partially dam-
aged (e.g., some broken bolts), the support damage is moder-

Draft manuscript – Copyright protected – Cai and Kaiser 2018


36 Rockburst phenomena and rockburst damage

ate. The damage is major if the support system is heavily


damaged and excessively deformed, and severe if the support
system fails to fulfill its intended function (modified from Pot-
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

vin 2009).
(3) If an excavation is unsupported or the support system is inef-
fective, the violence of the damage process is frequently de-
scribed by the estimated ejection velocity. Many velocity
magnifiers resulting from momentum transfer and failure
mechanisms may be involved and ejection velocity estimates
are therefore in general poor measures of the damage severity.
For this reason, the severity is best described by a measure of
seismic energy release, i.e., the magnitude of the seismic event
that is co-located with the damage.
Intensity of seismic event associated with rockburst
The intensity of seismic events associated with strainbursts is
affected by the burst volume, the local deviatoric stress, the rock
mass modulus and the local deformation potential (mine stiffness).
The related seismic event is typically characterized as an implo-
sive or crush event or it exhibits only a minor shear component
(weak double couple). The magnitudes of these seismic events are
relatively small with ML (local or Richter magnitude) mostly
ranging from 0 to 2. The signals of strainbursts are often hidden in
the signals from a larger, strainburst triggering (remote) seismic
fault-slip event.
When damage to an excavation is caused by a remote seismic
event, i.e., a fault-slip event, the rockburst damage may at least in
part be related to the intensity of the seismic event, i.e., the radiat-
ed energy due to slip at the seismic source.
For fault-slip events, the seismic energy release process is similar
to that of an earthquake but the released energy is typically much
smaller. The radiation patterns are focussed (not spherical) and the
seismic wave motion frequencies of the seismic signals associated
with rockbursts are higher than those of earthquakes as illustrated
in Figure 2-5. The high frequency ground motions attenuate more
quickly and thus have a smaller zone of influence. For fault-slip
events, the corner frequency and the logarithm of the energy (log
Es) are, amongst other factors, related to the radius of the seismic
source. Hence, smaller seismic events generally radiate less ener-
gy at higher corner frequencies.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 37

Very small seismic events or micro-seismic events are common in


underground mines. The wave motion frequencies of these events
are higher than those of fault-slip events and, while damaging the
rock mass, they do not cause rockbursts.
Audible by
Can be felt by human Only detectable by sensors
human ears

Low frequency High frequency


Audible acoustic
wave

Microseismic events
Rock bursts Acoustic emission Acoustic emission
Earthquakes (in underground
(in mines) (in rocks) (in metals & ceramics)
excavations)

Frequency spectrum

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
-1 0 1 2 3 4 5 6 7
10 10 10 10 10 10 10 10 10

Frequency (Hz)
Figure 2-5 Frequency spectrum for various types of seismic sources (Cai et
al. 2007).
Techniques developed in earthquake studies are frequently used to
characterize the intensity of rockbursts, i.e., to characterize the
energy release associated with the seismic event. The Richter or
Local magnitude ML and the Nuttli magnitude MN are used to
describe the size of a seismic event.
The most appropriate measure of the strength of a seismic source
is the radiated seismic energy Es and it forms the basis for a loga-
rithmic magnitude scale (Gutenberg and Richter (1954). The
Richter magnitude ML of earthquakes is determined from the
logarithm of the amplitude of recorded seismic waves. Because
the magnitude scale is logarithmic, a one unit higher magnitude on
the Richter scale is approximately equivalent to a 10 to 30-times1
higher energy release.
Because a routine estimation of seismic energy from waveforms is
demanding and the rate of seismic activity in mines does not
always allow for timely processing of seismic signals with proper
corrections for site effects, the energy estimates obtained by rout-

1
10 is based on Eq. (5.2a) in the CRBSPHB for events with magnitudes below 4.5 and 30
is based on Eq. (5.1) (Hanks and Kanamori, 1979) that is valid to large earthquakes.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


38 Rockburst phenomena and rockburst damage

ing data processing systems are often inaccurate. Event magnitude


values may differ significantly from one mine to another. For
large events, signals may be clipped and the magnitude may be
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

underestimated.
A simple solution for magnitude estimation in hard rock mines is
to utilize one of the seismic moment-based magnitude relations,
e.g., the Hanks and Kanamori (1979) relation. The resulting mag-
nitude is consistent with the Gutenberg and Richter empirical
relation between seismic energy and magnitude for intermediate to
larger earthquakes at constant apparent stress (Mendecki 2016). In
other words, the magnitude scale may be unreliable for smaller
rockbursts.
Most seismic events in underground mines are too small in magni-
tude to become rockbursts and the event intensity depends on the
rock failure mechanism. The largest fault-slip event recorded in
South Africa occurred on March 9, 2005, in the Klerksdorp district,
with an estimated magnitude of ML = 5.3 (Gibowicz 2009). This
event shook the nearby town of Stilfontein, causing serious dam-
age to several buildings and minor injuries to 58 people. Wide-
spread underground damage is typically associated with such large
seismic events.
Moderate to large rockbursts can be heard as loud rock fracturing
noise and shock waves can be felt both in the ground and in the air
in nearby excavations. Very small rockbursts can be heard as
popping sounds underground. Unless sensitive monitoring sensors
are used, human beings may not be able to feel the shock wave
generated by such small seismic events.

2.1.6 What has been done to deal with rockburst?


Considerable research effort at an international scale (Australia,
Canada, China, South Africa, USA, etc.) has been carried out to
understand the rockburst phenomenon. Eight conferences under
the theme of International Symposium on Rockbursts and Seis-
micity in Mines (RaSiM) have been held since 1982 and South
Africa, Canada, and Australia has undertaken major research
initiatives to study rockbursts and find engineering solutions to
eliminate and control rockburst hazards. Many papers have been
written and several influential books (Hedley 1992; Kaiser et al.
1996; Ortlepp 1997; Blake and Hedley 2003) have been published.
This reference book in many ways consolidates some of the works

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 39

of these studies and injects new thoughts and understandings.


Particularly the Engineering Seismology Group (ESG;
www.esgsolutions.com) and the Institute of Mine Seismology
(IMS; www.imseismology.org) have made major advances in
micro-seismic monitoring since the 1990s. Hundreds of seismic
monitoring systems are today in operation at mines and tunnel
construction sites around the world for operational decision-
making and for research purposes. From the waveforms recorded,
the time, location, radiated energy, seismic moment, and other
source parameters of a seismic event can be estimated. Monitoring
of seismic events in mines is a very useful tool for outlining poten-
tially hazardous ground conditions and it is assisting mine man-
agement in the implementation of effective re-entry policies
(Vallejos and McKinnon 2008). Advanced 3D numerical model-
ling (Beck et al. 2006) and visualization (Kaiser et al. 2005; Cai et

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
al. 2006) have been used to identify potentially hazardous areas
and assist in mine planning and design. Various rockburst support
products have been developed and used in mining operations to
help mitigate rockburst damage. Despite many outstanding efforts
made so far, rockbursts will continue to challenge the underground
construction and mining sectors and further research and im-
provements to the engineering process are needed.
With respect to support design, Kaiser and Cai (2013a, b) pointed
out that some of the commonly adopted principles are flawed and
need to be revised. More recently, Stacey (2016) suggested that it
is still not possible to follow conventional design approaches for
support design and proposed a risk-consequence approach where-
by decisions should be made based on quantifying risk measures.
This reference book attempts to summarize the current state-of-
the-art of rock support design and provides engineering principles
and quantitative means for support selection. This basic engineer-
ing approach is adopted here despite the valid concerns expressed
by Stacey (2016) who indicates that uncertainties in capacity and
demand, particularly in demand, may introduce ‘design indetermi-
nacy.’ It is the authors’ view that due diligence is required, that a
conventional design process must be followed with due respect for
the high levels of uncertainty when designing support for burst-
prone ground. There are many situations, particularly in grounds
with minor to intermediate burst severity, where both demands
and capacities can be established and ranges of variability as-
sessed with sufficient confidence.
Draft manuscript – Copyright protected – Cai and Kaiser 2018
40 Rockburst phenomena and rockburst damage

As pointed out in Chapter 3, support is the last line of defence and


other controls have to be used to mitigate rockburst risks. More
than in any other field of engineering, risks beyond those covered
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

in a systematic design evaluation process have to be determined


and covered as described by Stacey (2016). It is for this reason
that we recommend using the term ‘support selection’ rather than
‘support design’. Once an engineering evaluation has been com-
pleted following this reference book, a prudent engineer will have
to consider all other relevant aspects to arrive at viable solutions.

2.2 Types of rockbursts


As indicated in the previous section, there are three fundamental
rockburst types: fault-slip-, pillar- and strain-bursts. These rock-
bursts are either mining-induced due to static stress changes caus-
ing damage at or near the source, or caused by dynamic disturb-
ances; the latter is predominantly caused by dynamic stress pulses
or energy transfer resulting from a remote seismic event.
Ortlepp and Stacey (1994) further subdivided rockburst classes for
South African conditions into five types as summarized in Table
2-1. However, in a broad sense, these sub-classes fit the above
listed three types:
- Buckling type rockbursts typically occur in layered or foliated rock
and can be grouped with strainbursts. Buckling is driven by tangen-
tial, bedding parallel deformation and the released energy depends
on the stress level and this deformation. Therefore, a buckling rock-
burst is a strainburst. The energy release is related to the burst vol-
ume and is relatively small (0 < ML < 1.5).
- ‘Face crush’ is a particular form of pillar wall fracturing that is
frequently encountered in tabular ore bodies, i.e., in South Africa.
Pillar wall fracturing can, however, also be encountered in deep cav-
ing operations under high undercut loads on the drill and extraction
horizons. These ‘face crush’ bursts are strainbursts. The energy re-
lease is related to the volume of the failing rock and can reach mag-
nitudes as high as ML = 2.5.
- ‘Pillar burst’ is a rockburst where the entire pillar collapses either by
a sudden failure of the pillar core or by a progressive failure process
that propagates through the entire pillar. In situations where a pillar
failure leads to a domino effect or a pillar run, or if pillars are shear
loaded (Suorineni et al. 2011; Suorineni et al. 2014), magnitudes
higher than ML = 2.5 may be encountered. Pillar bursts are fortunate-
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 41

ly relatively rare. They cannot be prevented by rock support alone


and therefore are not the focus of this book.
- Shear ruptures can be grouped with fault-slip events as both are shear
failure processes occurring inside the rock mass. The difference be-
ing that a new fault is created or a fault is extended by breaking
through the rock mass. From a support design perspective, it is im-
portant to differentiate between fault-slip events that occur remote
from excavations and those that are near or intersect excavations (see
Section 2.2.3). The former cause dynamic disturbances whereas the
later also disrupt the rock mass structure and affect the excavation
failure modes.

Table 2-1 Classification of seismic event sources (modified from Ortlepp and
Stacey (1994)) with types of rockbursts identified in South African mines
Rockburst type First motion
Rockburst Magnitude

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
by Ortlepp and Postulated source mechanism from seismic
type ML
Stacey (1994) records
Usually
Superficial spalling with violent
Strainburst undetected, could -0.2 to 0
ejection of fragments
be implosive
Strainburst
Outward expulsion of large slabs
Probably
Buckling pre-existing parallel to surface of 0 to 1.5
implosive
opening
Sudden collapse of stope pillar, or Possible complex,
Pillar or face
Pillar burst violent expulsion of large volume implosive and 1.0 to 2.5
crush
of rock from tabular tunnel face shear
Violent propagation of shear Double-couple
Shear rupture 2.0 to 3.5
Fault-slip fracture through intact rock mass shear
burst Sudden movement along exiting Double-couple
Fault-slip 2.5 to 5.0
fault shear

In this reference book, the authors therefore only refer to fault-


slip-, pillar- and strain-bursts.

2.2.1 Strainburst
This type of rockburst was previously introduced as “a sudden and
violent failure of rock near an excavation boundary caused by
excessive straining of an un-fractured volume of rock.” Hence,
strainbursts occur when the stress near an excavation reaches the
peak strength of the unsupported or supported rock mass and the
rock suddenly fails by a combination of extension and shear frac-
tures.
This type of rock failure is either initiated by a stress change in-
duced by the tunnel advance or near-by mining, potentially trig-
gered by a dynamic stress pulse, or by a dynamic stress change
Draft manuscript – Copyright protected – Cai and Kaiser 2018
42 Rockburst phenomena and rockburst damage

due to seismic waves from a large remote seismic event. Hence, it


is important to differentiate between self-initiated, mining-induced,
seismically triggered, and dynamically loaded strainbursts. This
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

distinction is new, compared with the 1996 version of the CRB-


SHB, and is highly relevant from a support design perspective. A
detailed description of these strainburst types is given in the Ter-
minology section in Appendix A and the reader is referred to
Section 2.2.1 and two keynote papers published by the authors
(Kaiser and Cai 2013a, b).
Because of natural stress variability and rock mass strength heter-
ogeneity, the stress-to-strength ratio is rarely constant and strain-
bursts therefore are localized in areas where this ratio reaches a
critical value and the excavations are vulnerable.
Strainburst induced by tunnel advance and by mining
Strainbursts are the most common rockburst type in many under-
ground mines and they occur most often during the production
stage when mining changes the state of stress and strains the rock
mass near the excavation walls.
Two examples are shown in Figure 2-6. The first illustrates deformed
and damaged support that was able to ensure the integrity of the excava-
tion after a strainburst. Cracking and spalling of shotcrete (‘raining
shotcrete’) poses a safety risk unless mesh-over-shotcrete is added. The
second example presents a situation where the support below the first
row of bolting was not able to retain the stress-fractured ground, allow-
ing rock to be ejected.
Two conditions must be met for a strainburst to occur:
1) the rock mass must fail in a brittle manner, i.e., it must display a
high intrinsic brittleness, and
2) a high level of tangential stress must build up in the skin of the
excavation.
These two conditions largely define the strainburst potential (SBP)
and damage only occurs if the installed support system is ineffec-
tive.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 43

(a) (b)

Figure 2-6 Strainburst damage: (a) deformed ground support with cracked
shotcrete after a retained strainburst; (b) strainburst damage with rock ejection
near floor (photos courtesy: Grasberg and Kidd Creek Mine)

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Four additional factors define the strainburst severity (SBS):
1) the volume of rock that actually bursts2 which is called the burst
volume (Figure 2-2);
2) the energy imposed by the surrounding rock mass through forces
and deformations acting on the burst volume, i.e., a factor enhanc-
ing the relative brittleness3;
3) the energy consumed during the failure, i.e., the deformation of the
reinforced volume of fractured rock; and
4) the volume increase (bulking) due to stress fracturing of the burst
volume.
The larger the burst volume and the higher the energy imposed on
this volume is, the higher is the magnitude of the co-located seis-
mic event. When the burst volume fails, the tractions on the sur-
rounding ground are partially or fully removed and the burst vol-
ume is deformed, i.e., energy is released from the rock mass
surrounding the excavation and imposed on the burst volume. The
radiated seismic energy, resulting from a sudden removal of the
tractions acting on the rock surrounding the burst volume, pro-

2
The burst volume is less than the volume of failed rock (Figure 2-2). The latter includes
the stress-fractured rock surrounding the burst volume. The former is the volume of rock
that was not fractured before the burst.
3
Tarasov and Potvin (2013) discuss the concept of intrinsic and relative brittleness in the
context of laboratory tests. They also provide a discussion of limitations and deficiencies
of various indices proposed in the literature to describe the intrinsic brittleness of rocks.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


44 Rockburst phenomena and rockburst damage

vides a measure of the energy imposed on the burst volume and


therefore of the strainburst intensity. This energy input, by refer-
ence to energy release from a laboratory test frame, depends on
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

the loading system stiffness (LSS) or the mine stiffness in the field.
The lower the LSS or the softer the mine stiffness, the higher is the
energy input from the surrounding rock mass (Manouchehrian and
Cai 2016). Therefore, the higher the potential for deformation (or
deformation potential DP) upon the removal of the burst volume,
the higher is the energy imposed on the burst volume.
During tunnel advance in massive to moderately ground, without
the influence from nearby mining and intersecting geological
structures, the energy imposed from the surrounding rock mass is
relatively low (i.e., the LSS is high or the relative brittleness is
low). The strainburst severity is often dominated by the strength
and intrinsic brittleness of the rock blocks making up the rock
mass. In such situations, the ratio of uniaxial compressive strength
to uniaxial tensile strength (UCS/stens) provides a means to antici-
pate the spalling potential, and the stress level at the point of
failure is a simple indicator of the available energy (Diederichs
2014).
Simple relations as illustrated by Figure 2-7 are therefore suitable
to assess in a preliminary manner the strainburst potential and
severity, but only for conditions that are not affected by mining
and other factors that lower the loading system stiffness. This
figure therefore is only quantitatively applicable for relatively high
LSS-values.
Soft loading systems are created when multiple openings are
excavated close to each other, i.e., at high extraction ratios. Large
excavation sizes and unfavourable yield zone geometries also
lower the LSS. Geological weaknesses, such as faults or shears,
locally reduce the LSS by increasing the deformation (slip) poten-
tial. Hence, the energy release potential may locally be higher than
that without these structures. Because such structures may create
stress raisers and increase the degree of freedom, larger volumes
of rock may be involved in the deformation and failure process.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 45

Strainburst severity SBS


S
SB
nd
Pa
h SB
Hig

Strainburst potential SBP

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Figure 2-7 Strainburst potential and severity in massive, homogeneous ground
without mining and geological influences (modified after Diederichs 2014).
Various energy sinks including rock fractures, friction, heat and
support system deformation consume this imposed energy. If these
energy sinks collectively balance the energy input, equilibrium is
re-established and the excavation remains stable (there is no sup-
port failure or rock ejection). Otherwise, parts of the rock mass
and support system will be ejected. The ejection velocity is merely
in part related to the energy input as only the difference between
energy input and consumption is available to eject rock (for more
detail see Section 2.4.5 or Volume II on strainbursts). The ‘visible’
portion of the energy release, as reflected in the ejection of rock
blocks or parts of the support, depends on the amount of energy
consumed in the burst volume and the reinforced fractured rock
surrounding the burst volume.
In mining- or stress-induced strainbursts, the rockburst damage is
related to the local mine stiffness and the post-peak behaviour of
the supported rock and is not directly related to the intensity of the
associated mine seismicity.
However, even if equilibrium is re-established, the burst volume
increases due to rock mass bulking during the fracturing process
and this imposes radial deformations as schematically illustrated
by the displacement d in Figure 2-2. The entire bulking defor-
mation has to be directed toward the excavation, i.e., in the radial
direction, because the fractured rock can only move toward the
Draft manuscript – Copyright protected – Cai and Kaiser 2018
46 Rockburst phenomena and rockburst damage

excavation. If this bulking occurs suddenly, the bulking process


imposes a dynamic radial displacement pulse. Equilibrium is re-
established and the excavation is stable if the supported rock mass
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

and the support system are able to resist this deformation (without
support or rock ejection). Otherwise, part of the rock mass and the
support system will be ejected. The ejection velocity is not related
to energy input, because some of it is consumed by the tangential
deformation imposed on the burst volume and its bulking factor as
well as the speed of failure (for more details see Volume II on
strainbursts). Therefore, the ‘visible’ portion of the strainburst, as
reflected in rock or support deformation or ejection, will largely
depend on the bulking characteristics of the fractured rock and the
failure speed of the burst volume.
Strainbursts may be violent when much energy is available or the
bulking rate is high, or rather mundane if all input energy is con-
sumed during the rock fracturing process and the bulking process
is gradual.
In both cases, the fractured rock mass will eventually occupy a
larger volume than before fracturing. This bulking deformation
constitutes the static or dynamic deformation demand on the sup-
port. Hence, a key principle of a deformation-based support design
(see Volume II) is that, no matter how violent a rock mass fails,
the support has to be able to survive the deformation imposed by
the bulking of the stress-fractured rock volume.
Seismically triggered strainbursts
Unstable failure processes, such as strainbursts, can easily be
triggered by even small disturbances. For this reason, it is im-
portant to differentiate between failures caused by seismic energy
input and failures that are triggered by seismicity only. A seismi-
cally triggered strainburst means that a mining-induced strain-
burst is triggered by a remote seismic event. There are at least two
events present in this case. The remote event constitutes the pri-
mary seismic event whereas the seismic event co-located with the
strainburst damage is a secondary event. However, the damage is
not related to the intensity of the remote primary seismic event,
which serves only as the trigger of the strainburst.
The likelihood and the timing of seismically triggered strainbursts
may be related to the mine seismicity. Larger seismic events add
seismic disturbances at greater distances from a seismic source
and thus are more likely to trigger a strainburst or even multiple
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 47

strainbursts. The location, not the intensity of seismically triggered


strainbursts, may therefore be affected by mine seismicity.
In a similar fashion, strainbursts can be triggered by production
blasts. In fact, when an excavation is severely damaged by a pro-
duction blast, it is very likely that the excavation was already
vulnerable to strainbursting. The blast then serves as the trigger
rather than the supplier of the damage-causing energy. If a damage
location is extremely close to a large blast, by adding energy the
blast may cause dynamically loaded strainbursts.
Dynamically loaded strainbursts
As described in the Terminology section (Appendix A of Volume
I), a strainburst is classified as a dynamically loaded strainburst
when a remote event causes a substantial dynamic stress incre-
ment near the damage location. This dynamic stress increment
may deepen the depth of failure and the ground motions may add

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
energy to the failing rock. Unless a remote seismic event adds
substantial stress, the damage is dominated by the pre-existing
conditions at the strainburst location.
Part of the damage causing stress or energy will stem from the
remote seismic event that causes the dynamically loaded strain-
burst. Hence, the conventional wisdom of relating damage direct-
ly and entirely to the remote seismic source’s characteristics could
be, and is frequently, flawed. For support design against strain-
bursting, it must be respected that the ground condition and load-
ing system stiffness at the damage location often controls the
demand on the support.

2.2.2 Pillar bursts


A pillar burst, as the name implies, is defined as a violent failure
in the pillar core, e.g., by shear rupture, or a complete pillar col-
lapse. The seismic source and rock failure locations are co-located
in the confined core of the pillar. An example of a pillar burst is
shown in Figure 2-8.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


48 Rockburst phenomena and rockburst damage
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Figure 2-8 An example of a pillar burst with pillar on the left side of the image
(Hedley 1992).

Violent pillar wall bursting often occurs as precursors of pillar


burst. They are strainbursts and should not be classified as pillar
bursts. In this case, it is the excavation wall and not the pillar that
has failed.
A pillar burst may be triggered by a remote seismic event although
the damage is dominated by the energy released from the failing
rock in the pillar core and the surrounding rock mass. In a low
loading system stiffness environment such as room and pillar
mining at a high extraction ratio and under high overburden stress,
pillar burst can become extremely violent and may lead to pillar
runs (sequential pillar bursts).
The released energy from the volume of failed rock and the sur-
rounding rock mass is larger than in strainbursts causing seismic
events with magnitudes in the range of 2 < ML < 3 (if shear rupture
occurs in highly confined pillar cores). The seismic event magni-
tude associated with pillar bursts depends on the mining depth, the
rock mass strength and modulus, the local mine stiffness, and the
pillar and stope geometry. Unstable and violent pillar failures
occur when the local mine stiffness is soft compared with the post-
peak stiffness of the pillar (Salamon 1970). In Ontario mines in
Canada the highest recorded pillar burst event magnitude was
reported as MN = 3.5 or ML = 3.0 (Kaiser et al. 1996).
The most disastrous consequence of pillar bursting presents a
scenario where the initial burst leads to a domino effect with
sequential pillar bursting as experienced at Coalbrook Mine in
South Africa in 1960 (Salamon 1999) and at Solvay Mine, USA,
in 1995 (Board et al. 2007). The former claimed 437 lives and the

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 49

latter was recorded as a magnitude 5.1 seismic event when rough-


ly a 1 by 2 km2 section of the mine collapsed.
If pillars are more competent than the surrounding rock mass,
shear ruptures may also form in the hanging wall or footwall. A
failure may be more violent as a larger volume of failing rock is
involved and ruptures occur in highly confined conditions.

2.2.3 Fault-slip rockbursts


Because geological structures such as faults affect both the mechanisms
of energy release and the excavation damage processes, it is necessary to
distinguish between remote (non-intersecting) fault-slip rockbursts and
rockbursts involving structural slip of intersecting structures.
Remote fault-slip rockbursts
When a fault-slip event occurs remote from an excavation, rock-

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
burst damage is caused by the energy radiated from a dynamically
slipping, pre-existing fault, fault zone or from a newly generated
shear rupture. A critically stressed fault, with shear stresses reach-
ing its shear strength, can slip, particularly when the degree of
freedom is changed as it is intersected by a mine opening or by the
yield zone surrounding a mining area. The strength of faults is a
function of the normal stress, the coefficient of friction of the fault
surface, its waviness or dilation characteristics, and the strength of
rock bridges, which are called asperities in earthquake terminolo-
gy. A fault may therefore slip when the shear strength is reduced
due to a drop in the clamping stress. The introduction of water (e.g.
from drilling) may also lower the shear strength of a fault or fault
zones.
Damage can also be caused by shear rupture through massive or
moderately jointed rock masses (Figure 2-9).
Such shear rupture bursts have been observed in South African
mines (Ortlepp 1997, 2000), with magnitudes exceeding ML = 3.5.
Ortlepp (1997) strongly advocates shear rupture as one of the most
important source mechanisms for major rockbursts, and fractures
such as the one shown in Figure 2-9 are called ‘Ortlepp shears.’
Ortlepp (1997) captured the image below from an area that previ-
ously experienced major fault-slip events after it was mined
through. It must be noted that shear fractures, as observed in South
African mines, are created in a rather soft (constant overburden
pressure) mining system. Even though shear ruptures can occur in

Draft manuscript – Copyright protected – Cai and Kaiser 2018


50 Rockburst phenomena and rockburst damage

confined pillar cores, lessons learned from South Africa may not
always be applicable to non-reef type mining environments.
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Figure 2-9 Rupture No. 18 (Ortlepp 1997); ruler pointing at shear rupture.

Recent work by Bewick (2013) and (Bewick et al. 2014a, b;


Bewick et al. 2014c), shows that the characteristics of a shear
rupture zone are not only a function of the rock or rock mass
properties but also depend on the prevailing boundary or confining
conditions under which the rupture zone is created. In soft envi-
ronments, shear rupture is facilitated by the ease of dilation. An
example from a numerical model (Bewick et al. 2014b) is shown
in Figure 2-10 for low and high normal or confining stress condi-
tions.
Shear rupture zones are localized at a low normal stress (Figure
2-10a). They become increasingly complex with en echelon frac-
tures making up a wider shear zone when created at higher normal
or confining stress as shown in Figure 2-10b. It should be noted
that the ultimate rupture zone is not necessarily planar; it is em-
bedded in a relatively wide rock mass damage zone, which may be
composed of associated en echelon shears, as they by themselves
are short faults. In such a case, the focal plane solutions of seismic
sub-events will not be aligned with the direction of the rupture
zone during the early stages of the shear rupture creation. Only
once a shear rupture is fully developed and has undergone signifi-
cant deformation would focal plane solutions be aligned with the
slipping direction.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 51

As a general guide, shear ruptures created at low normal stress


display more post-peak brittleness and thus could be more prone
to violent failure (Bewick et al. 2014a and b). Due to the brittle-
ness of such shear ruptures, they are typically associated with a
single large seismic event. At a high normal stress, repeated seis-
mic events may be expected when the rupture zone is developing
as the failure process is generating multiple en echelon shears to
eventually create a shear rupture zone.
In stiff environments where dilation is prevented or constrained, the coupling
between the normal and shear stresses (stress-path) under constant normal
stiffness boundary conditions leads to a different rupture development process.
The process is displacement- rather than stress-controlled and occurs pre-peak.
Due to the displacement-controlled failure process, ongoing mining-induced
deformation is needed to fully form shear ruptures in stiff environments. As
illustrated by

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Figure 2-11, the eventual shear is localized but also associated
with en echelon fractures. Sub-figure (b) illustrates the shear
rupture pattern with 40% more shear deformation than in sub-
figure (a).

(a) (b)

Figure 2-10 Shear rupture created in a soft, dilative environment under


relatively low to high normal stress: (a) 5 MPa and (b) 25 MPa (Bewick et
al. 2014b).
As a general guide, shear ruptures created at low normal stress
display more post-peak brittleness and thus could be more prone
to violent failure (Bewick et al. 2014a,b). Due to the brittleness of
such shear ruptures, they are typically associated with a single
large seismic event. At high normal stress, repeated seismic events
may be expected when the rupture zone is developing as the fail-

Draft manuscript – Copyright protected – Cai and Kaiser 2018


52 Rockburst phenomena and rockburst damage

ure process is generating multiple en echelon shears to eventually


create the shear rupture zone.
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

(a) (b)

Figure 2-11 Shear rupture created in a stiff, non-dilative and highly confined
environment and with increasing shear deformation (shear displacement in (b) is
40% greater than in (a); Bewick et al. 2014a).
Most importantly, large stress shear stress adjustments (or large
seismic events) can be expected long before the shear rupture has
fully formed, i.e., even before the peak strength of the rupture
zone has been reached. Repeated larger stress drops or shear stress
oscillations are encountered due to cohesion loss during the rup-
ture zone formation. Hence, repeated seismic activity is indicative
of the formation of a shear rupture zone in a stiff environment.
Fault-slip and shear rupture rockbursts typically occur in deep
mines when the extraction ratio is high and large closures are
allowed to persist over large mining volumes. Associated fault-
slip events may release large amounts of seismic energy, coming
from the instantaneous partial relaxation of the elastic strain ener-
gy stored in a volume of highly stressed rock surrounding the slip
or rupture area, and radiate seismic energy in the form of com-
pressive (P) and shear (S) waves. These ground vibrations or
ground motions may trigger strainbursts or pillar bursts, cause
dynamically loaded strainburst or shakedowns, or eject insuffi-
ciently supported rock by energy or momentum transfer to broken
rocks.
In a strict sense, dynamically loaded strainbursts are fault-slip
bursts as they are associated with a remote seismic event resulting
from fault slip or shear rupture. The distinction between a dynam-
ically loaded strainburst and a fault-slip burst is somewhat fuzzy
as there is a gradation from one to the other. In a dynamically
loaded strainburst, most of the energy comes from the rock mass
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 53

surrounding the burst volume, whereas most of the energy causing


damage from a fault-slip event originates from the rock mass
surrounding the fault or rupture zone. Consequently, it is possible
or even likely that both energy sources contribute to the damage
process. If the damage is primarily caused by the energy radiated
from a dynamically slippage on a pre-existing fault, fault zone, or
from a newly generated shear rupture, it should be classified as a
fault-slip burst. If the energy is primarily due to strainbursting, it
is to be classified as a dynamically loaded strainburst.
It is often difficult to identify the pre-dominant cause of failure or
damage process from observations of the damaged excavation and
support. Material size and shape of the failure zone (notch) may
give some indication of the predominant process. Finer rock frag-
ments and a notch-shaped failure zone might be indicative of
strainbursting, whereas larger rock blocks might suggest that some

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
structural controls with energy or momentum transfers had con-
tributed to the failure.
Rockbursts involving structural slip on structures intersecting tunnels
Geological structures such as faults intersecting excavations are
locally less restrained and free to slip, i.e., slippage or relative
movements between opposite sides of the fault in the rock mass is
facilitated by the excavation. This increases the displacement
potential and thus the associated energy release. It may also facili-
tate shear rupture propagation at the ends of short geological
structures or rupture of asperities within the structures. Related,
often devastating, rockbursts should more appropriately be called
‘structural slip rockbursts’ or ‘structurally controlled strainbursts’
(Diederichs 2014), rather than fault-slip bursts, as the damage is
dominated by the ability of the local geological structure to slip
into the excavation. Support design principles applicable to remote
fault-slip rockburst are not directly transferable to structural slip
rockbursts. The later are often the result of a combination of struc-
turally assisted or structurally controlled strainbursts and seismi-
cally induced falls of ground (shakedowns).
As discussed in more detail in Volume II, geologic structures can
influence the stress level SL and the loading system stiffness LSS.
Intersecting faults and geological structures therefore enhance the
rockburst potential and the rockburst severity. Weakness planes,
faults or fault zones also tend to increase the vulnerability of
excavations to strainbursting. They inevitably add more degrees of

Draft manuscript – Copyright protected – Cai and Kaiser 2018


54 Rockburst phenomena and rockburst damage

freedom for rock block movement and increase the potential and
extent of structurally controlled failures (unraveling). This also
renders locations in a mine with intersecting geological structures
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

more prone to shakedown.


Structures that are sub-parallel and close to excavations increase
the local SL, promote stress fracturing and once exposed behave
like intersecting structures. They are particularly hazardous be-
cause it is often difficult to detect such sub-parallel structures,
posing a hidden risk with high rockburst potential and severity.

2.3 Mine seismicity causing dynamic disturbances


Mining-induced stress changes may trigger strainbursts or cause
fault-slip events, radiating seismic ground motions or displace-
ment waves. These dynamic disturbances impose differential
displacements and hence dynamic stress waves. The ground mo-
tions are characterized by the wave length (frequency) and ampli-
tude, and are typically described by the peak ground displacement
(PGD), velocity (PGV) or acceleration (PGA), and the corner
frequency (f0). The maximum magnitude of the associated stress
wave can be related to the PGV, the shear or compression wave
propagation speed (cs or cp ), and the density (ρ) of the medium.
Furthermore, the stress waves can be amplified by stress wave
reflections, e.g., at excavation boundaries, or refractions, e.g., at
geological contacts, and by stress relaxations that depend on the
mine geometry.
Even though dynamic disturbances in the form of stress waves are
related the PGV, this does not imply that the ground velocity
constitutes the predominant demand that needs to be considered
for support design. To the contrary, in most cases, it is the acceler-
ation that may shakedown a marginally stable volume of (stress-
fractured) rock or it is the stress increment that brings a vulnerable
excavation to the failure point in the form of a seismically induced
strainburst.
Volume III presents a detailed assessment of seismic risk and
hazard for ground control planning together with means to derive
inputs for ground-motion based support design. Here, we present a
brief overview of some basic ground motion principles that help to
understand the potential impact of dynamic disturbances on differ-
ent excavation instability modes and on support selection.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 55

2.3.1 Seismic source characteristics


Event magnitude
The local magnitude scale ML is used in this guide to characterize
the intensity of seismic sources. It is related to the seismic moment
M0 [in GN×m] and the dynamic stress drop Ds [in MPa]:
M L = log M 0 + log Ds - (2.0 ± 0.15). (2-1)

A detailed discussion of the limitations and use of event magni-


tudes for support design is presented in Volume III. This includes
descriptions on how to select a design event and its location.
Seismic energy radiation
For a fault-slip source or a double couple (DC) source model, the
seismic wave or energy radiation pattern is schematically illustrat-
ed by Figure 2-12. It shows that the P- and S-waves are focused at

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
45° to each other and because the P-wave magnitude is typically
much smaller than the S-wave magnitude, the highest dynamic
disturbances are encountered in the slip or s-wave direction (or
perpendicular to it).

P-wave

S-wave

P-wave and S-wave


radiation pattern

Figure 2-12 Schematic radiation patterns for a DC source model (modified after
Aki and Richards (2002)).
Because the orientation of a slip event is not a priori known, it is
necessary to assume, for support design purposes, that the maxi-
mum disturbance could be encountered in any direction. For this
reason, it is meaningful to assume a ‘spherical radiation pattern’
and to scale the intensity of the dynamic disturbance as a function
of the distance from the seismic source to obtain the ‘design
ground motion’. This does not mean that the actual radiation
pattern is spherical. Much smaller dynamic disturbances will be

Draft manuscript – Copyright protected – Cai and Kaiser 2018


56 Rockburst phenomena and rockburst damage

measured at many locations surrounding a seismic source. This is


discussed in more detail in Volume III.
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

2.3.2 Ground velocity for support design


McGarr (1984) proposed a scaling law that can be written in non-
logarithmic form as:
*
M 0a (2-2)
PGV = C * ,
R

where M0 is in GN×m, R in m, and a* and C* are empirical con-


stants.
The constants a* and C* may vary between mine sites, but Kaiser
et al. (1996), based on the analysis of seismic data from a global
database, found that a* can be fixed at a* = 0.5 for many mines.
Average C*-values, obtained from log (R×PGV) vs. log (M0Ds)
plots, range from 0.04 to 0.13 (for stress drops of Ds = 1 to 10
MPa). Mean PGV values are of little use for a design as, by defini-
tion, 50% of locations would experience higher PGV. For design
purposes it is necessary to define an upper-bound limit for a de-
sired confidence (e.g., 90 or 95% confidence). Kaiser et al. (1996)
recommended C*-values ranging from 0.17 to 0.64 (for Ds = 1 to
10 MPa and 90 to 95% confidence).
In terms of event magnitude, assuming a* = 0.5, the ground motion
for support design PGVD can be obtained from:

10( M L +(1.5±0.15)) (2-3)


PGVD = C * .
R

This design scaling law should only be applied to openings in the


far-field (Kaiser and Maloney 1997), at distances greater than Rmin:
1/ 3
æM ö
R min » 1.5 ç 0 ÷ . (2-4)
è Ds ø

As a general guide, scaling laws should not be used to obtain PGV


at locations closer than 10 m from the seismic source (for strong
seismic events, not closer than 30 to 50 m).
Near surface amplification of dynamic disturbances
Stress waves can be reflected at excavation boundaries and this, in
combination with other factors, can lead to magnifications of
ground motions at or near excavation walls. As discussed in Vol-

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 57

ume III, the actual PGV near excavation boundaries can locally be
as much as 2 to 5-times higher than those obtained by a scaling
law. Because ground motion measurements, used to obtain C*,
from log (R×PGV) vs. log (M0 Ds) graphs, originated mostly from
sensors embedded far from excavation boundaries, C* does not
capture the effects of wave reflections and amplifications near
excavations. Hence, some adjustments to the C*-values quoted in
Kaiser et al. (1996) may be justified for support design purposes.
From Eqs. (2-2) or (2-3), it follows that PGV is proportional to C*
and could locally be as much as (n = 2 to 5)-times higher than that
given by the scaling law. However, the average magnification
factor within the supported rock volume must be lower. Unfortu-
nately, no systematic study is available to assess the average mag-
nification in supported ground volumes. Dynamic wave models
provide some insight (see Volume III), but the complexity of

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
unknown source models and related radiation patterns, 3D exca-
vating geometries as well as unknown attenuation characteristics
for stress-fractured and yielded ground prevent accurate forecast-
ing of ground motions for design purposes.
In situations where there is field evidence of excavation or support
damage that might be attributed to wave magnification, it might be
justified to adopt higher C*-values than those proposed by Kaiser
et al. (1996). Because the entire volume of supported rock must be
affected by the magnification, it is not reasonable to assume 2 to
5-times higher values, which might lead to an uneconomical de-
sign. It is suggested that between 1.5 and 2-times higher C* values
can be adopted in such cases. For support design purposes, it may
therefore be advisable to adopt C*-values ranging from 0.25 to 1.3
(for Ds = 1 to 10 MPa and 90 to 95% confidence) in areas with a
high stress wave interaction. Such high C*-values may lead to high
support demands and result in costly solutions. Hence, it is rec-
ommended that field verifications be used to justify such extreme
design parameters.

2.3.3 Ground acceleration for support design


For sinusoidal-shaped waveforms, the peak particle acceleration
PGA is frequency dependent and can be related to the PGV by:
PGA = 2p × f × PGV (2-5)

Draft manuscript – Copyright protected – Cai and Kaiser 2018


58 Rockburst phenomena and rockburst damage

where the frequency f = v/l, i.e., the phase speed of the wave v
divided by the wavelength l.
The wavelength must be significantly longer than the extent of a
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

potential rockfall to simultaneously and uni-directionally acceler-


ate the entire rockfall volume. Consequently, dominant frequen-
cies of 10 to 30 Hz are most critical for potentially unstable wedg-
es or volumes of well retained broken rock and typical mine
opening dimensions. This frequency range has been adopted in
this book for the analysis of seismically induced falls of ground.
In naturally blocky or stress-fractured ground with block sizes in
the decimeter-range, however, higher frequencies of 1 to 10 kHz
can accelerate individual blocks or rock fragments and this may
initiate unraveling processes. Because such frequencies lead to
extremely high dynamic forces, it is not reasonable to assume high
frequencies for overall support loading. It does though explain
why individual blocks of rock may experience very high accelera-
tion forces that locally can impact retention systems such as mesh
and shotcrete.

2.3.4 Dynamic ground stress


A seismic event or blast may add an increment of dynamic stress
and disturb the in-situ and mining-induced stress field. This dy-
namic stress increment may trigger strainbursts or increase the
depth of stress-fractured ground.
The dynamic stress pulse of the shear wave modifies the in-situ
state of stress by principal stress increments Δσ1d = + cs ρ PGVs
and Δσ3d = - cs ρ PGVs, 4 where PGVs is the peak ground velocity
of the shear wave. For example, a ground motion wave with PGVs
= 0.3 m/s would cause a far-field principal stress change of about
2.5 MPa. This stress wave repeatedly loads the rock mass, though
with decreasing stress intensity as the ground motion decays from
its peak value.
As the stress wave passes an excavation, it alters the stress con-
centration near the underground opening. For a circular tunnel in
elastic ground, the maximum dynamic stress increment at the
excavation wall will oscillate as described by:

4
The principal stress difference is twice as large as the dynamic stress increment due to
sign reversal. Because the ground motion from the p-wave is normally much smaller than
that from the shear wave only the latter is considered here.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 59

Δσdmax = ± 4 cs ρ PGVs (2-6)

For a low frequency ground motion wave with PGVs = 0.3 m/s,
the maximum pseudo-static tangential stress change will be as
high as ±10 MPa. This will repeatedly increase the maximum
stress concentration near the excavation wall and simultaneously
relax the minimum stress around the excavation. As a conse-
quence, the depth of stress fracturing may increase in one direc-
tion and relaxed ground may unravel in the other.
Details are discussed in Volume III and implications for strain-
bursting and stress fracturing are described in Volume II.

2.4 Rockburst damage mechanisms


Understanding all possible excavation damage mechanisms by
rockbursts is critical for deriving strategies to eliminate or mitigate

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
rockburst hazards, i.e., to work out tactics to reduce the burst
potential or to select rockburst resistant support.
Rockburst damage mechanisms are invariably a consequence of
the static and dynamic demands on the rock mass and the support
system reaching and exceeding the capacity of the supported rock
mass.
Excavations loaded by dynamic disturbances fundamentally show
the same failure modes as excavations that fail under static loading
except that the dynamic factors modify the static forces, stresses
and deformations as well as the deformation rates.
In mining, excavations are increasingly strained as the extraction
ratio increases or other factors lead to mining-induced stress
changes with related deformations. When affected by dynamic
disturbances, described in more detail in Volume II, a highly
stressed (or strained) excavation may approach the failure point.
Image (1) in Figure 2-13 shows an example of tangential straining
of the drift wall due to floor heave or roof sag.
When excessively strained and stressed or relaxed, excavations
may experience one of three basic dynamic failure modes (Kaiser
et al. 1996):
- Shakedown with stand-up time reductions (Figure 2-13: Image (2)).
This failure mode is dominated by the rock quality, excavation span,
amongst other factors, and dynamic acceleration forces from a re-
mote seismic event or other dynamic disturbances;

Draft manuscript – Copyright protected – Cai and Kaiser 2018


60 Rockburst phenomena and rockburst damage

- Static stress fracturing or strainbursting due to tangential straining


with or without rock ejection (Figure 2-13: Image (3)). This failure
mode is dominated by stored strain energy, the loading system stiff-
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

ness LSS and the in-situ stress field, causing tangential deformations
in the direction indicated by large arrows. Associated rock mass
bulking causes large static and dynamic deformations near the exca-
vations, which are defined by the depth of failure, the bulking factor
and the mining-induced tangential strain; and
- Rock ejection by momentum or energy transfer from remote seismic
sources (Figure 2-13: Image (4)) or from high bulking deformation
rates during strainbursts (Figure 2-13: Image (3)). This failure mode
is dominated by energy transmitted from remote seismic sources for
fault-slip events and, most importantly, by the fracture rate due to
strainbursting.

Figure 2-13 Illustration of mining-induced tangential straining of drift walls due


to floor heave or roof sag (1) and three dynamic failure modes near excavations:
(2) shakedown due to acceleration forces from a remote seismic event;
(3) static stress fracturing or strainbursting due to tangential straining; and
(4) rock ejection by momentum or energy transfer from a remote seismic source
or due to high bulking deformation rates during strainburst.
All failure modes can be assisted by dynamic disturbances from
remote seismic events; e.g., shakedowns by temporary accelera-
tions, stress fracturing by stress pulses, and ejection by energy
transfer mechanisms. However, as indicated previously, excava-
tions may be more or less vulnerable and damage can also be
caused by strainbursting as a result of static overstress. This is
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 61

confirmed by the damage data presented in Figure 2-14 from


Morissette et al. (2012). Data points identified by them as strain-
bursts are circles in this modified figure.
The trend lines5 in Figure 2-14 were added by the authors of this
book for data points on either side of ML = 2.0. For events with
magnitudes less than 2.0, the distance to damage locations is
essentially independent of the seismic event magnitude with dis-
tances ranging from < 10 m to about 100 m (average 25 to 40 m;
few > 100 m) from the seismic source. No damage was observed
between 15 and 100 m (about 30 to 40 m on average) from the
seismic source. This suggests that the damage was caused by
strainbursts triggered by remote seismic events and that the dam-
age severity for ML < 2 is dominated by the energy stored in the
rock mass surrounding the burst volume rather than the radiated
energy from the remote seismic events. The trend line for all

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
strainbursts identified by Morissette et al. (2012) up to ML = 2.6
(circled in this figure) is only slightly inclined, similarly indicating
a range from 25 to 40 m for 0 < ML < 2.
For the data with ML > 2, identified by Morissette et al. (2012) as
mostly associated with damage triggered or caused by remote
fault-slip events, the trend lines for ‘moderate’ and ‘important’
damages (trend lines D’ and C’ in Figure 2-14) are steeper and the
distance to the damage location therefore increases with increasing
seismic event magnitude. As will be shown later (Figure 2-25), the
trend line for ML > 2 follows the anticipated trend of constant
PGV-lines.

5
These trend lines are statistically not equivalent; as a matter of fact, some are poorly
correlated (e.g., the line for ‘no damage’ with only one point for ML < 2 and six points for
ML > 2).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


62 Rockburst phenomena and rockburst damage

1000

ML < 2 ML > 2
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

A: No damage

B: Moderate (ML < 2)


D': Trend line
B': Moderate (ML > 2)
Distance to source (m)

C: Important (ML < 2)


100
C': Important (ML > 2)
D: Trend line C': Trend line
D: Major (all)
B': Trend line
All strainbursts Strainbursts
Trend line Linear (A: No damage)
A: Trend line
(No damage)
Expon. (B: Moderate (ML < 2))

Expon. (B': Moderate (ML > 2))


10 C: Trend line
Expon. (C: Important (ML < 2))

Expon. (C': Important (ML > 2))

Expon. (D: Major (all))

Expon. (Strainbursts)

1
-1 -0.5 0 0.5 1 1.5 2 2.5 3 3.5 4
Richter magnitude ML

Figure 2-14 Magnitude–distance relation encountered at Creighton mine in


Canada for four damage indices (Morissette et al. 2012). Also shown are expo-
nential trend lines for data on either side of ML = 2 and for data identified as
strainbursts (large red circles).
This dataset suggests that excavation damage is dominated by the
excavation vulnerability and the local energy release potential. For
events with low magnitudes (< 2.0), failure can be triggered at
relatively large distances if the excavation is vulnerable, and high
damage severities can be found at large distances where local
conditions facilitate much energy release. For events with higher
magnitudes, the dynamic disturbance tends to cause damage at
increasingly greater distances away from the seismic source, i.e.,
as far as 200 m from a source with ML = 3.5. However, there is no
clear indication in this dataset that the damage severity (damage
rating) is higher closer to the source. As a matter of fact, ‘no dam-
age’ (diamond) was observed at ML > 2 closer to the source than
locations with damage severities of ‘moderate’ to ‘important’.
In summary, this dataset confirms that the primary role of a re-
mote seismic event is to trigger the failure process and that the
damage severity and distance to the damage location from the
seismic trigger event is largely related to the excavation vulnera-
bility and the local available energy release.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 63

2.4.1 Seismically induced falls of ground


Marginally stable ground (e.g., wedges) and stress-fractured rocks
can be brought to failure by seismic shaking or by a temporary
increase in gravitational forces (Figure 2-13: Image (2)). Seismi-
cally induced rockfalls are caused by (low frequency) seismic
waves from relatively large and remote seismic events that shake
the entire volume of a potentially unstable mass of rock. The
incoming seismic wave accelerates a volume of rock that was
previously stable under static loading conditions, adding dynamic
forces that may trigger a fall and potentially overcome the capaci-
ty of the support system, particularly if the static factor of safety is
marginal.
There is also a possibility that the first incoming stress wave dy-
namically fractures a volume of rock and subsequent vibrations
accelerate this fractured rock, initiating an unravelling or shake-

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
down failure. Seismically induced rockfalls frequently occur at
intersections where the span is large and the confinement of the
roof rock is low. Examples of a seismically induced rockfall and
an unraveling shakedown failure are presented in Figure 2-15.
Seismically induced rockfalls are particularly common in deep
mines when rock fracturing occurs in the backs due to high sub-
horizontal principal field stresses. However, heavy stress-damage
in the walls, e.g., under an undercut in a caving operation, may
also facilitate seismically induced falls if the effective span is
enlarged by deep fracturing in the walls.
For seismically induced rockfalls, the damage causing energy is
primarily derived from gravitational forces. The dynamic disturb-
ance from a remote seismic event often only triggers a fall by
breaking rock bridges, thereby reducing the rock mass’s self-
supporting capacity, or by adding a temporary acceleration 𝑎𝑎
�⃗, thus
increasing the demand from m×g to m×(g + 𝑎𝑎 �⃗), where g is the
gravitational acceleration. Seismic waves also temporarily alter
the normal stress conditions on geological structures, with a dy-
namic unclamping effect leading to a temporarily lower shear
strength (Kaiser et al. 1996). Hence, seismically induced rockfalls
can be caused by either one of these two factors or a combination
of both.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


64 Rockburst phenomena and rockburst damage
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

(a) (b)
Figure 2-15 Examples of seismically induced shakedown failures: (a) fall of
ground with large blocks of rock and pulled-out bolts; and (b) unraveling of
stress-fractured ground.

2.4.2 Rock mass bulking


When the tangential stress (shown in Figure 2-13: Image (3) by two
arrows) reaches the strength or load bearing capacity of the reinforced
rock mass near an excavation, the supported rock mass will fail by grad-
ual stress fracturing or sudden strainbursting.
In brittle rocks, extension fracture initiation and propagation,
combined with shear along pre-existing weaknesses (e.g., joints)
or newly created inclined fractures, lead to a disintegration of the
rock mass and this is associated with rock mass bulking. In hard
rock, this bulking process is primarily a result of geometric in-
compatibilities between rock blocks or fragments as fractured rock
blocks do not fit into the same space as before failure. This can
lead to large volume increases. For example, blasted rocks will
have a volumetric swell or a bulking factor of 35 to 40%.
Due to the boundary conditions surrounding the burst volume, the
volumetric change has to be accommodated in the radial direction.
This inward movement can be described by a radial or linear
bulking factor (BF). The actual volume increase varies considera-
bly, and the linear bulking factor BF is generally large (up to 25%)
for unconfined highly fragmented rocks and smaller for rock that
is confined and remains in place even after being fractured (BF <
5 to 10%). If the rock mass is well-supported, this bulking can be
restrained to smaller values. However, because of the geometric
non-fit, this unidirectional deformation, perpendicular to the wall,
is much larger than that predicted by standard constitutive dilation
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 65

models (Detournay 1986; Alejano and Alonso 2005; Zhao and Cai
2010) for yielding rock masses.
The effect of bulking can be visually observed as illustrated by the
examples shown in Figure 2-16. It can be measured by extensome-
ters or estimated from convergence or laser scan surveys. In high-
ly stressed grounds, rock mass bulking cannot be prevented but
can be controlled or at least partially suppressed by effective
support measures (see an example in Figure 2-16c). Heavy rock
mass bulking between tendons is often observed when the bolt
spacing is high or the retention system too flexible.

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
(a) (b) (c)
Figure 2-16 Examples of rock mass bulking after rockbursts: (a) bulking that
caused failure of some bolts and bulging of mesh; (b) same but with bulking
behind mesh-reinforced shotcrete, and (c) bulking constrained by a yielding
support system with shotcrete, mesh and straps.
Another indication of rock mass bulking is the frequently ob-
served floor-heave associated with rockbursts. Because floors are
typically unsupported, once the rock mass is fractured, its volume
increases freely, leading to high bulking factors in the floor.
For rock support design, it is necessary to anticipate and control
the depth of fracturing and the volume change due to bulking (for
details see Volume II).

2.4.3 Rock mass buckling


In laminated or foliated ground, thin rock layers may buckle rather
than spall or bulk in a combination of extension and shear fractur-
ing.
Euler-type buckling of a laterally unconstrained layer occurs when
the slenderness ratio exceeds a critical value, i.e., Le/r > 89, where
Le is the effective length and r is the radius of gyration (for rectan-
gular cross-section: r = 0.289t, where t is lamination thickness).
Three hinged buckling, as shown in Figure 2-17, occurs when the
Draft manuscript – Copyright protected – Cai and Kaiser 2018
66 Rockburst phenomena and rockburst damage

exposed slab length is > 25.7t, e.g., at a length > 500 mm for t =
20 mm.
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Figure 2-17 Buckling of foliated rock (image is about 1 m wide) (Courtesy:


Cuiabá mine – AngloGold Ashanti).
The lateral displacement imposed on each layer causes a geomet-
rically controlled inward movement that decreases with distance
from the wall because the buckling length and the lateral defor-
mation decrease toward the apex of the buckling notch. A sche-
matic model, assuming a three-hinge equivalent for the buckled
slabs, with an imposed displacement of 1 mm at the tunnel surface
and at each end of the slab is presented in Figure 2-19 (every
second slab interface is shown). The lateral displacement is set to
zero at the depth where buckling is prevented (at 250 mm where L
= Le = 500 mm for this example).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 67

500

d = 1 mm or 0.2%
400
Notch height and deflection (mm)

300

200

100

-100
0 500 1000
Buckling length (mm)
Figure 2-18 Schematic buckling pattern for a 1 m wide notch with 20 mm thick

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
slabs (every second slab boundary is shown) deformed at each end by 1 mm for
0.2% tangential strain.
If an extensometer were installed in the centre, the schematic
displacement profile presented in Figure 2-19 for this 1 m wide,
0.5 m deep buckling notch would be recorded. For a layer thick-
ness of 20 mm, buckling is prevented at a depth of 250 mm where
the buckling length is insufficient and crushing starts to dominate.
As is evident from the buckling image and profile, the bulking
factor (change in radial length per unit length) in this buckling
ground can be rather high. The overall bulking factor BF depends
on the tangential strain (0.2% in this case); BF is equal to 10% for
the illustrative example. Most importantly, the largest displace-
ment jump, approximately 10 mm in this case, is observed at the
deepest point where buckling is still possible but the buckling
length is the shortest. Strain localization with associated straining
of bolts therefore would occur at the interface between buckling
and crushing ground.
The above presented figures are intended as schematic illustrations
of the buckling and related bulking process. They are based on the
assumption that there is only geometric buckling. In reality, how-
ever, there will be compressive failure at the hinges; the overall
inward displacement will be less. Nevertheless, these illustrative
examples show that buckling grounds can impose large and non-
uniform displacements on the support (both the bolts and the
retaining components).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


68 Rockburst phenomena and rockburst damage
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Figure 2-19 Schematic extensometer profile through a buckling zone generated


by considering geometric buckling without crushing of layers.
The effective buckling length can be increased by the application
of radial support pressures. There are two limits: a lower limit to
stop the buckling process and a higher limit to prevent buckling
and enforce a crushing mode in the lamination slabs. For the ex-
ample presented above, radial pressures in the order of 0.05 MPa
are sufficient to balance the critical buckling stresses. Thigh mesh
can therefore stabilize the buckling process. However, radial
pressures of about 0.2 MPa would be required to prevent buckling
for the case simulated above. Even tightly bolted, relatively thick
layers of fibre-reinforced shotcrete cannot prevent the initiation of
the buckling process but it would limit the depth of buckling and
drastically reduce the bulking displacements.
The bulking profile presented in Figure 2-19 as a full line is only
valid if the notch forms at the excavation surface as shown in the
photo of Figure 2-17. If deep-seated buckling occurs, the con-
finements or the stresses perpendicular to the laminations remain
low and deep decompression zones develop. Karampinos et al.
(2015) studied squeezing ground conditions at La Ronde mine in
Canada and demonstrated, using 2D and 3D discontinuum models
(UDEC and 3DEC), that deep decompression zones can develop
((Figure 2-20a) with high bulking gradients near the wall (dis-
placement contours in Figure 2-20b with overlay of the schematic
notch bulking chart of Figure 2-19)).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 69

Inward displacement [mm]


-35

-30

-25

-20


-15

-10

-5

0
0 Distance from wall [mm]
100 200

(a) (b)
Figure 2-20 (a) Minor principal stress distribution with lateral decompression
zones (white contour lines) and (b) associated displacement patterns from 3DEC
modelling showing shear and buckling failures of laminated ground (modified
from Karampinos et al. 2015).

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
As the decompression zone deepens, laminations near the wall
will buckle together and bulking is limited in this zone as illustrat-
ed by the dashed line in Figure 2-19. Rockbolts will be strained
primarily in the zone of high dilation (i.e., high convergence gra-
dient). In other words, while the decompression zone deepens,
lockbolts will be strained further inside the buckling rock mass.
From a practical perspective, it is important to differentiate be-
tween bulking caused by buckling or by spalling and shear failure.
For buckling ground, surface pressure can reduce the depth of
buckling and rock reinforcement can increase the slab thickness
and consequently the length of rock slabs that are prone to buck-
ling. In burst-prone ground, however, thicker slabs will store more
energy and therefore may promote more violent buckling-type
failures.

2.4.4 Rock mass bulking enhanced by shear rupture


If the slenderness ratio is less than the critical value, i.e., Le/r < 89,
buckling is prevented and shear failure has to occur at the yield
strength of a slab. In brittle rock, this occurs by kink-band for-
mation at the interface between the stress-fractured and undis-
turbed ground as illustrated in Figure 2-21a (white dashed arrow).
This process can occur in a controlled matter, if the loading sys-
tem stiffness LSS is high and deformations are induced gradually,
e.g., as during tunnel advance. If the LSS is sufficiently low and
the stress level is high enough to rapidly propagate failure, the

Draft manuscript – Copyright protected – Cai and Kaiser 2018


70 Rockburst phenomena and rockburst damage

kink-band formation may lead to shear rupture as illustrated in


Section 2.2.3 (e.g., Figure 2-10a). Because of the weakest link
concept, this shear slip or rupture typically occurs only on one side
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

of a notch but may trigger shear ruptures on opposite sides of a


tunnel as highlighted by the black dashed arrows in Figure 2-21b.
The wall slab will ‘kick-out’ (out of the wall; into the excavation)
at the shear slip or rupture location where the displacement is the
highest.

(a) (b)
Figure 2-21 (a) Surface spalling with kink-bands causing shear at the interface
between the stress-fractured and undisturbed rock mass (white dashed arrow);
and (b) shear localization (x = shear) at edge of stress-fractured zone (o =
tension) during ‘notch’ formation in continuum model (RS2) (black arrows
show locations of shear failure).
This shear rupture process increases the bulking of the stress-
fractured ground (Figure 2-21a) and focuses the associated dis-
placements at one side of the notch. If this occurs suddenly, the
displacement velocity is the highest near the rupture plane and this
may lead to local rock ejection with enhanced ejection velocities if
the stress-fractured ground is not retained.
By application of a similar model as used for the buckling analysis
but assuming one-sided slip :EHG@:`BG<EBG>= LA>:KHK KNIMNK>
plane, the ‘kick-out’ displacement for a flat wall slab can be esti-
mated. This is shown for wall slabs of 1, 2, and 3 m length in
Figure 2-22. For these slab lengths, the ‘kick-out’ displacements
are 100 mm at 1, 0.25, and 0.1% wall strain, respectively.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 71

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Figure 2-22 Slab displacement at the ‘kick-out’ location for three slab or notch
heights.

2.4.5 Rock ejection


Rock ejection can be caused by strainbursts (e.g., Figure 2-13:
Image (3)), by a remote seismic event through dynamic energy or
momentum transfer (Figure 2-13: Image (4)), or by a combination
of both. When a rock mass suddenly fractures, violent bulking can
impose high deformation velocities and thus eject rock from the
burst volume as well as broken rock between the burst volume and
the tunnel boundary. An ejected rock block may further transfer
energy by momentum transfer to smaller fragments and magnify
these ejection velocities 6. An example of rock ejection is shown in
Figure 2-23.
When a large (fault-slip) seismic event occurs relatively close to
an excavation, the radiated seismic waves reaching an under-
ground opening may accelerate rock blocks to a velocity propor-
tional to the PGV of the shear stress wave. These ground motions
may be amplified near excavation boundaries due to the interac-
tion of incoming seismic waves with the tunnel (Wang and Cai

6
When assessing rock ejection, it is important to consider the effect of momentum
transfer from large to small blocks. For a mass ratio of 100, the ejection velocity differs
by a factor of 10; thus, a 0.1 ´ 1 ´ 1 m3 slab moving at 0.3 m/s can eject a block of 1 dm3
at 3 m/s.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


72 Rockburst phenomena and rockburst damage

2015; Wang and Cai 2016). Durrheim (2012), for example, found
that ground motions at the surface of excavations in South African
mines were amplified 4 to 10-times. Ejected rocks may travel at
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

velocities in excess of 3 m/s and up to 12 m/s as estimated by


Ortlepp and Stacey (1994), however, the upper end of this ejection
velocity range cannot be explained by direct energy transfer
mechanism from seismic sources alone (Footnote 7).

Figure 2-23 Example of excavation damage with rock ejection. Large rock and
shotcrete pieces were ejected 2 to 3 m away from the pillar nose (photo courtesy:
Kidd Creek Mine).
If the ground motion is sufficiently high, and the rock near the
surface is already stress fractured and inadequately supported,
individual blocks or fragments may break away from the deeper,
more massive and confined rock mass and eject into the opening.
This rock ejection could theoretically be caused exclusively by
energy transfer from a remote seismic source, although it is more
likely that the rock ejected by the energy coming from energy
stored in the failing rock and the surrounding rock mass.
It is therefore unlikely that rock ejection is caused by energy
transfer from the remote seismic source alone as other sources of
energy are simultaneously released. Hence, it is not advisable to
design support based on an energy transfer criterion alone. As
previously discussed in conjunction with Figure 2-14, there does
not seem to be any evidence in the rockburst dataset that the dam-
age severity and the distance between damage and source location

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 73

is directly related to the intensity or magnitude of the damage


triggering seismic event (at least for ML < 2).
If rock ejection occurs during a strainburst, the ejection velocity is
not directly to the energy transmitted by a remote seismic event7.
It is in large parts related to the locally stored or releasable strain
energy (minus energy consumed during rock fracturing). Conse-
quently, it is incorrect to assume that the observed ejection veloci-
ty is directly related to the magnitude of the seismic source (e.g.,
fault-slip event) or the associated PGV. It is likely that a signifi-
cant component of the ejection causing energy stems from the
highly stressed rock near the excavation, particularly when the
failing rock is located in a soft loading system.

These considerations lead the authors to revise the support design


approach previously presented in the CRBSHB by including

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
considerations of strainburst energy release and associated dis-
placement demands on the support. Unfortunately, it is rather
difficult to accurately estimate the energy released from strain-
bursts. Means of quantifying the releasable stored strain energy
will be discussed in Volume II.
Furthermore, it is extremely difficult if not impossible to predict
whether rock will fail violently as part of the fracturing and bulk-
ing process but it is possible to anticipate the rockburst related
displacements. It is much easier to anticipate the displacement
rather than the energy imposed on the support by violently failing
rock. For this reason, a deformation-based support design ap-
proach is introduced in this volume and then quantified in Volume
II.(Mendecki 2013)
In practice, it is often sufficient to recognize that large parts of the
damaging energy stems from the burst volume and that the related
bulking displacement causes the support damage rather than mo-
mentum transfer from a remote seismic event. Identifying rock
ejection as a possible strainburst damage mechanism is an im-
portant step in assessing the damage potential and for designing
effective rock support systems.

concept for guiding damage assessment and support design.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


74 Rockburst phenomena and rockburst damage

2.5 Factors influencing rockburst damage


Because of the variability in many factors that influence rockburst
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

damage, the vulnerability of an excavation to rockburst damage is


highly variable, as is discussed in more detail in Volume II. Figure
2-24 summarizes influence factors and groups them into four
categories: geotechnical, geology, mining, and seismicity. As a
consequence, the severity of damage, defined by the depth of
failure and the lateral extent of the damage or the tonnage of dis-
placed rocks, can vary greatly over small distances in a mine.

Geotechnical Geology Mining Seismicity


•In-situ stress •Rock type •Mining-induced static •Seismically induced
(magnitude and •Foliation and stresses (excavation dynamic stresses
stress ratio) spans, mining and ground
bedding method)
•Rock strength •Geological motions
•Local mine stiffness
•Rock mass structures (dykes, (extraction ratio, rock •Event magnitude
quality/joint fabric faults, and shears) mass modulus) •Distance to seismic
•Rock mass •Excavation sequence source
brittleness (stress-path) •Source mechanism,
•Production rate e.g., fault slip or
(blasting, cave strainburst
loading)
•Rate of seismic
•Destressing and energy release
hydro-fracturing
•Effectiveness of
installed rock support
•Backfill

Figure 2-24 Main factors influencing rockburst damage potential and severity.

2.5.1 Geotechnical factors


The in-situ stress increases with depth and that is why rockburst
problems increase as mining migrates to deeper levels. A high
deviatoric stress, due to a large differential between the maximum
and the minimum principal stresses, generally induces high tan-
gential stresses near excavation boundaries. Hence, the in-situ
stress magnitude and stress ratio defines the overall stress level in
a mining block and this stress is modified by excavation geome-
tries. The local stress level combined with the rock mass strength
and deformation behaviour control the amount of strain energy
that can be stored in the rock and eventually be released during
failure. Strong and stiff rocks can build up high stress and store a
large amount of strain energy. The peak strength and the post-peak
stiffness of the rock mass (intrinsic stiffness) and the unloading
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 75

stiffness of the surrounding rock (relative stiffness) determine the


amount of releasable energy. Joint fabrics weaken the self-
supporting capability of the rock mass and this elevates the risk
for shakedowns and rock ejection.

2.5.2 Geology
Geological features, such as dykes, modify the mining-induced
stresses and often produce local stress raisers. Faults promote
stress concentrations and facilitate rock mass failure; they also
alter the local mine system stiffness. Brittle hard rock tends to
accumulate strain energy and it has the potential to fail violently
with little warning. Faults that intersect excavations enhance the
rock deformation, reduce the local loading system stiffness, and
thus increase the amount of releasable energy. Foliations and
beddings induce anisotropy in stress and strength and may alter

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
failure mechanisms.

2.5.3 Mining
Mining disturbs the in-situ stress field and the related stress
changes lead to stress relaxation and stress concentration zones
throughout a mine. Excavations alter the mine stiffness, generally
leading to a lower mine stiffness at a higher extraction ratio. When
the extraction ratio is high (typically > 80%), the remaining rock
will be highly stressed and the mine system stiffness will be low,
and as a consequence the likelihood of rockbursting is increased.
For example, the risk of rockbursting is high when recovering sill
pillars.
The excavation sequence determines the stress path and the rock
failure process (Kaiser et al. 2001; Cai 2008). Centre-out mining
sequences in open stoping and transverse cut-and-fill stoping are
widely practiced to reduce the rockburst risk. Stoping sequence
retreating from faults or shears results in a more even seismic
energy release (Hedley 1992), reducing the risk of large rockbursts.
The adopted mining method can therefore influence the rockburst
damage potential and severity. For example, most mines that
convert from cut-and-fill to long-hole stoping experience larger
rockbursts (Blake and Hedley 2003). At the Lucky Friday mine in
the USA, where a traditional overhand cut-and-fill mining method
was used until 1986, a switch over to the Lucky Friday Underhand
Longwall (LFUL) method was introduced in mid-1987 to manage
the rockburst risk (Jenkins et al. 2006).
Draft manuscript – Copyright protected – Cai and Kaiser 2018
76 Rockburst phenomena and rockburst damage

High tunnel advance or stope extraction rates increase the risk of a


rockburst as the released energy increases and not enough time is
available to allow the ground to settle down and reach a new
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

equilibrium. It was demonstrated by (Salamon 1983) that the


amount of released energy can be drastically reduced if the num-
ber of mining steps is increased.
A properly designed and installed rock support system can limit
the extent of damage and control the rockburst damage severity.
Excavations supported by a combination of reinforcements and
yielding bolts along with strong surface retention (e.g., mesh-
reinforced shotcrete) may effectively maintain the stability of
openings (e.g., Figure 2-16). Rock support is the last line of de-
fence and should be used in combination with other rockburst
control measures to minimize the rockburst potential and severity.
Backfilling of mined-out stopes is practiced in some burst-prone
mines to control mine-wide deformations and thus to reduce rock-
burst risk. In addition, some mines conduct destress blasting to
fracture rock ahead of the tunnel or stope face to reduce the strain-
burst risk during mine development. More recently hydro-
fracturing has been introduced as an effective means of seismic
energy release control (Araneda and Sougarret 2008; Catalan et al.
2017).

2.5.4 Seismicity
As discussed in Section 2.2, a strainburst can be triggered by a
remote seismic event or occur without any dynamic disturbances.
However, severe rockburst damage is often associated with a
fault-slip seismic event because the associated dynamic disturb-
ance causes dynamically loaded strainbursts. The larger a seismic
event is, the further away from the event the dynamic disturbance
can be felt and the wider the strainburst trigger zone will be. On
the other hand, the closer the seismic source to an opening is, the
greater are the dynamic disturbances, stress pulses or ground
motions, and therefore the rockburst potential. For this reason,
large seismic events such as fault-slip events increase the area for
potential rockburst damage and the severity of damage. Although
it is generally true that the damage increases with the event magni-
tude, there are many exceptions as other factors (listed above and
in Figure 2-24) influence the severity of rockburst damage. The
data presented earlier (Figure 2-14), however, suggests that this is
only valid for events with ML > 2. A large seismic event may also
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 77

cause ‘aftershocks’, i.e., secondary events such as strainbursts, and


this may cause rockburst damages at multiple locations.

2.5.5 Combination of factors


Some of the factors listed in Figure 2-24 determine the intensity of
dynamic loading and others determine the site response. Therefore,
rockburst damage is governed by a combination of these factors
that define the intensity of a dynamic disturbance and the vulnera-
bility of an excavation to rockburst damage.
For example, the combination of high stress, brittle rock, high
extraction ratio, and faults renders a mine highly burst-prone.
Excavations are at risk when large seismic events occur, and when
the installed rock support system is ineffective, severe damage is
inevitable.
Durrheim et al. (1998) found that the source mechanism is often

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
controlled by the mine layout, and by regional structures such as
faults, shears and dykes, whereas local rock conditions and sup-
port systems primarily influence the location and severity of dam-
age.
In mines experiencing rockbursts, it is often observed that severe
damage is encountered at one place while immediately adjacent
locations remain undamaged. This can be attributed amongst other
factors to:
- highly non-uniform radiation patterns with highest ground motions
from the shear waves in the slip direction of a fault-slip source (and
zero or very low PGVs :M`MHBM);
- reflections, refractions, amplification and shielding of stress waves
on geological boundaries or underground excavations; and
- rock mass heterogeneity with rapid changing stress and strength
conditions.
The interaction of stress waves with excavations and support is
very complex and additional research is required to address re-
maining deficiency in rock support design resulting from a lack of
understanding dynamic rock mass damage processes. Additional
discussions on this topic are presented in Volumes II and III.

2.5.6 Dealing with rockburst damage


As illustrated by Figure 2-24, many influencing factors such as
mining-induced stresses, mine stiffness, excavation sequence,
production rates, effectiveness of installed rock support, and back-
Draft manuscript – Copyright protected – Cai and Kaiser 2018
78 Rockburst phenomena and rockburst damage

fill placement fall into the mining activity category. These factors
can be controlled by mine design and by mining operations. They
often provide the most effective means to reduce rockburst risk. A
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

sound rockburst mitigation strategy does not rely on a burst-


resistant rock support alone.

2.6 Rockburst damage severity


The severity of damage can be viewed from three perspectives: the
volume of failed or displaced rock, the degree of support damage, and
the violence of the energy release in terms of impact or ejection velocity
(if rock is displaced). For reporting purposes, the volume of displaced
rock is typically used as a measure of damage severity. For example, a
rockburst with more than five tonnes of displaced rock has to be reported
to the Ministry of Labour in Ontario, Canada. For support design purpos-
es, however, the volume of failed rock, the associated bulking defor-
mations, and the support loading rate or impact velocity have to be
considered to properly define the potential damage severity that has to be
mitigated by the support.

2.6.1 Classification of damage severity


Kaiser et al. (1996) suggested that the rockburst damage severity
is best characterized by the depth and lateral extent of the rock
involved in the failure process around an opening. A rockburst
damage is categorized into one of three severity levels: minor,
moderate, and major or severe, as shown schematically by Figure
2-4. Detailed descriptions of the severity levels can be found in
Kaiser et al. (1996).
This classification of damage severity considers primarily the
depth of failure. It is applicable to opening sizes commonly found
in underground hard rock mines, ranging in width and height from
3 and 6 m. The severity levels, however, depend on many other
factors (Kaiser et al. 1996) including the rock mass quality, sup-
port effectiveness, local mine stiffness, geological structures,
opening size and orientation, and intensity of seismicity.
These severity levels are based on three primary considerations:
- if half an annulus fails and the failed rock bulks by 30%, about 90,
60 and 20% of the excavation profile, respectively, will remain open
for the three severity levels; and

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 79

- if the fractured rock bulks in the damage zone but does not collapse
or unravel, the bulking deformation at BF = 10 to 20% reaches be-
tween 25 to 150 mm of radial deformation and this fractured rock
will be highly susceptible to unraveling between tendons.
At the minor severity level, full tunnel access is still available after
the burst and the deformation is small enough not to compromise
the entire support system. On the other hand, at the major damage
severity level, most support systems are severely damaged and
access is prevented.
Furthermore, if a rock mass fails in a violent manner to a depth of
failure of more than 0.75 m, high deformation rates have to be
expected due to sudden rock mass bulking. Under these conditions,
there is a high potential for individual rock component failure and
at least partial rock ejection.
The classification of damage severity by Blake and Hedley (2003)

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
also considers the volume of displaced rock. The damage severity
is classified as light, medium, and heavy when the displaced rock
is less than 10 t, between 10 and 50 t, and greater than 50 t, re-
spectively. The above two criteria are comparable, if the damage
area ranges from 15 m2 for light to 25 m2 for medium severity.
The rockburst damage scale (Table 2-2) presented by Potvin
(2009), which evolved from Kaiser et al. (1992) and Blake and
Hedley (2003), considers damage to both the rock mass and the
support system. These two items are related as an effective rock
support system can reduce or limit the amount of displaced rock.
For Table 2-2 to be of practical value for support design, the area
of damage needs to be specified for each class. Based on the au-
thors’ experience, the area is less than 10 m2 for R3, less than 50
m2 for R4, and larger than 50 m2 for R5. This information has
been integrated into Table 2-2.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


80 Rockburst phenomena and rockburst damage

Table 2-2 Rockburst damage scale (modifed from Potvin 2009)


Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Rockburst Rock mass Damage Rock support damage


damage scale damage surface area
R1 No damage, 0 No damage
minor, loose
R2 Minor dam- < 1 m2 Support system is loaded, loose in mesh, plates de-
age, less than formed, shotcrete cracked
1 t displaced
R3 1–10 t dis- < 10 m2 Some broken bolts, mesh bulged, shotcrete fractured
placed
R4 10–100 t 10 to 50 m2 Major damage to support system; retention capacity
displaced severely compromised
R5 100+ t dis- > 50 m2 Complete failure of support system
placed

2.6.2 Assessment of rockburst damage severity


In this section, it will be demonstrated that the commonly assumed
dependence of damage severity on seismic event intensity (magni-
tude) is rarely valid. This provides the justification for a need to
overhaul commonly adopted energy-based support design princi-
ples.
Heal et al. (2006) compiled a rockburst catalogue from various
underground metalliferous hard rock mines in Australia and Cana-
da containing 83 case histories with 254 damage locations. 43% of
the damage data points are located in stope accesses or cross-cuts
and 27% in stopes or ore drives, where mining-induced stresses
are high. A further 26% are located at intersections of various
excavation types and the remaining 4% are located in declines or
ramps, where mining-induced stresses are relatively low and rock
support is relatively intense because of the permanence of these
structures. 65% of the damage occurred in the backs (roof) and 35%
in the sidewalls. This suggests that two thirds of the damages are
gravity-dominated and thus are seismically triggered or seismical-
ly loaded falls of ground. The remaining one third are likely seis-
mically triggered or dynamically loaded strainbursts.
With respect to the support performance, 19% of the damage
resulted in ‘naked tendons’, meaning that lockbolts were left
protruding from the stable and more competent rock mass, and
that broken rock between bolts had unravelled or was ejected.
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 81

Interestingly, only 5% of the rock mass failure extended beyond


the embedment depth of the holding support elements (typically
2.4 m long). This indicates that the depth of failure was limited to
a relatively shallow portion of the underground openings and that
the retaining systems were often inadequate. Hence, rockburst
support systems must be chosen to manage the failure process in a
relatively shallow skin of fractured rock (typically < 2 m deep)
around an excavation.
Because any combination of the influence factors listed in Figure
2-24 affects rockburst damage, it is tempting to try to simplify
matters and relate damage to a single, presumably dominating
parameter. For example, Hedley (1992) presented a scaling law8
and used the resulting peak ground velocity PGV, as a function of
seismic event magnitude and distance to the seismic event source,
to empirically classify rockburst damage types and severity into

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
four classes:
- no damage (PGV < 50 mm/s),
- falls of loose rock (50 < PGV < 300 mm/s);
- falls of ground9 (300 < PGV < 600 mm/s); and
- severe damage (PGV > 600 mm/s).
It is important to note that these limits refer to falls of rock and
therefore may or should not be applied to strainbursting ground.
This approach defines two thresholds for seismically induced falls
of rock (falls of loose rock, i.e., small volumes, and falls of ground,
i.e., large volumes). Because the damage-causing PGV threshold
depends on the static factor of safety FSs, it seems meaningful to
define two trigger limit ranges for falls of ground: 50 to 300 mm/s
trigger of falls of loose rock when the excavation is largely stable,
even during rock bursts, and 300 to 600 mm/s when excavations
are marginally stable in advance of a dynamic disturbance and
thus can produce falls of ground. Considering an overlap of the
two ranges, one would expect vulnerable excavations to experi-

8
This scaling law was later revised by Kaiser et al. (1996) based on a large dataset from
Creighton mine in Canada. The general applicability with an adjustment for the near-field
range was recently confirmed by Mendecki (2013).
9
By making the distinction between falls of loose rock and falls of ground, Hedley (1992)
differentiated between statically stable conditions with FSs > 1 producing mostly loose
rock and marginally stable conditions with FSs ~ 1, conditions that can lead to falls of
ground.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


82 Rockburst phenomena and rockburst damage

ence moderate to severe damage by falls of ground when a PGV


range of about 200 to 600 mm/s is reached or exceeded. Excava-
tions that are effectively supported and sub-critically stressed
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

might be stable even at these PGV thresholds (refer to Section


2.6.3).
This is supported by an examination of the data presented in Fig-
ure 2-14. For comparison purposes, PGV limits based on Kaiser et
al. (1996) with parameters calibrated for Creighton mine in Cana-
da have been superimposed to produce Figure 2-25. From this
figure, as is discussed next, it follows that the use of PGV limits
alone is too simplistic. It is not suitable to relate ground motion
limits to rockburst damage without considering other dominating
factors such as the vulnerability of an excavation or the static
factor of safety and strainburst potential at a time when the exca-
vation is affected by a dynamic disturbance.
1000

ML < 2 ML > 2

D': Trend line


Distance to source (m)

100
D: Trend line C': Trend line

B': Trend line


All strainbursts
Trend line
A: Trend line
(No damage)

10 C: Trend line

A: No damage
B: Moderate (ML < 2)
B': Moderate (ML > 2)
C: Important (ML < 2)
C': Important (ML > 2)
D: Major (all)
Strainbursts
1
-1 -0.5 0 0.5 1 1.5 2 2.5 3 3.5 4

Richter magnitude ML

Figure 2-25 Reproduction of Figure 2-14 with damage thresholds for PGVest =
50, 100, 300, and 600 mm/s (terminate at near-field limit of applicability).
Creighton mine parameters with C* = 0.1 for mean conditions are applied
(Kaiser et al. 1996).
More than 25 damage locations of the dataset from Creighton
mine (Morissette et al. 2012) should not have shown any damage

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 83

(data above the R*50 line10); in reality most of them (> 90%)
exhibited strainbursts. Furthermore, only 1 of 6 undamaged loca-
tions actually fell into the ‘no damage zone’ with PGV < 50 mm/s
(see also Section 2.6.3 on ground motion tolerance of stable exca-
vations). All but 6 of the remaining damage cases would have to
be classified as falls of loose rock or falls of ground (data between
the R*50 and R*600 lines; note that about 65% of these are identi-
fied as strainbursts).
As indicated above, the vulnerability of an excavation to failure
and strainbursting renders this single parameter (PGV) approach
as deficient, particularly when the stress level in advance of a
dynamic disturbance is ignored. Because equally vulnerable exca-
vations should not experience damage for R > R*50, it follows,
according to Hedley (1992), that all 26 damage cases above the
R*50 line must be self-initiated or triggered strainbursts or seismi-

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
cally induced falls of ground. 22 of the 26 cases were indeed
identified by Morissette et al. (2012) as strainbursts. Furthermore,
as discussed earlier, the flat trend lines for the ‘moderate’ and the
‘important’ damage level at ML < 2 indicate that there are equal
percentages of a given damage level closer or further from a given
seismic source and that there is no dependence of the distance to
the damage location or the source intensity.
It is interesting, but not statistically relevant11, that for ML > 2 the
trend of the distance R from a seismic source to a damage appears
to be essentially parallel to the constant PGV lines for PGV = 150
to 200 mm/s. This supports the view that the primary impact of
larger seismic events is to expand the zone of influence where
strainbursts or seismic shakedown can be triggered. It does not
mean that the damage severity is related to the magnitude of the
seismic event.
Realizing that not all excavations are equally vulnerable, Heal et
al. (2006) proposed to assess the rockburst damage severity by
introducing an Excavation Vulnerability Potential index EVP (for
more detail see Volume II). Because the approach adopted by
Heal et al. (2006) is heavily biased by a static damage initiation
and depth of failure factor (based on Kaiser et al. (1996)), this

10
‘R*xx’ refers to the distance R from the seismic source where xx represents the ground
motion PGV = xx m/s).
11
Note there is a lack of data for ML > 2.7 and R < 100 m.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


84 Rockburst phenomena and rockburst damage

approach is not applicable to conditions where falls of ground,


making up two thirds of Heal’s dataset, are caused by seismic
shaking.
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Furthermore, for their forensic analyses to establish threshold


values, both Hedley (1992) and Heal et al. (2006) used scaling
laws to estimate the PGV at the damage locations. For forensic
analyses, this approach is fundamentally flawed as the actual
ground motions causing the damage are, in most instances, much
smaller than those predicted by an applicable design scaling law
(Kaiser and Cai 2013a). In other words, the PGVs at the damage
locations are likely less and possibly much less than indicated by
the threshold limits (R*50 to R*600) plotted in Figure 2-25. Syn-
thetic ground motion modelling tools would have to be adopted to
obtain realistic PGV at the damage locations (see Volume III).
In summary, most of the data points correspond to situations
where strainbursting causes the damage or constitutes part of the
damage process; i.e., conditions where most of the damage caus-
ing energy stems from the rock mass surrounding the burst volume
and not from the damage triggering seismic event. Morissette et al.
(2012) therefore, correctly classified most of the damages for ML <
2, and some of the damages at ML > 2, as being caused by strain-
bursts. It follows that the commonly assumed dependence of
damage severity on seismic event intensity (magnitude) is rarely
valid and that the commonly adopted energy-based support design
principles are rarely applicable.

2.6.3 Ground motion tolerance of stable excavations


It is often observed that excavations close to seismic sources
remain undamaged while others further away experience minor to
severe damage. As a matter of fact, it is frequently observed that
conditions along a tunnel rapidly change from ‘severe’ to ‘no
damage’. This is illustrated in Figure 2-26 (modified Figure 2-25)
by the six ‘no damage’ data points from Morissette et al. (2012)
and 15 additional ‘no damage’ cases from a mine where these
seismic events caused severe rockburst damage at ‘equally sup-
ported’ excavations and at much larger distances from the seismic
source (locations are not shown). At these locations, the excava-
tions must have been less (or not) vulnerable to rockburst damage,
either because the rock mass near the excavation was not burst-

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 85

prone or the support proved to be more effective. They survived


estimated ground motions12 between PGVest = 50 to > 600 m/s.
Three of them survived even in the ‘near field’ of the respective
seismic events with ML ranging from 1.6 to 3.0. It follows that
these excavations were not vulnerable and ground motions emitted
by the seismic events were insufficient to trigger strainbursts or
cause other modes of failure (e.g., seismically induced falls of
ground) at these locations. All but one of these ‘no damage’ loca-
tions are between 8 to 50 m (plus or minus source location error)
from seismic events with ML = 1 to 3 where minor to most severe
damage would be expected based on a ground motion criteria that
only considers PGV as a design parameter.
.
1000

ML < 2 ML > 2

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
D': Trend line
A: No damage
Distance to source (m)

B: Moderate (ML < 2)


100
B': Moderate (ML > 2)
D: Trend line C': Trend line

B': Trend line C: Important (ML < 2)


All strainbursts
Trend line C': Important (ML > 2)
A: Trend line
(No damage) D: Major (all)

Strainbursts
10 C: Trend line
Shortest distance to excavation
50 mm/s without damage

100 mm/s

300 mm/s
1
-1 -0.5 0 0.5 1 1.5 2 2.5 3 3.5 4

Richter magnitude ML

Figure 2-26 Reproduction of Figure 2-25 with 15 seismic events plotted at the
nearest supported excavations that were not damage (Data from Mine A over a
14-year period).
It follows again that the commonly assumed dependence of dam-
age severity on seismic event intensity (magnitude) is not suffi-

12
The term ‘estimated ground motion’ or PGVest represents the ground motion calculated
by assuming radial radiation from a point source, i.e., by using a scaling law or ground
motion equation. Due to radiation patterns, the actual ground motion is likely less than
the PGVest.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


86 Rockburst phenomena and rockburst damage

cient to anticipate the rockburst damage severity and that the


commonly adopted energy-based support design principles must
be overhauled. Furthermore, it can be concluded that ground
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

motion hazard criteria (or hazard maps based on PGV alone) are
not sufficient to identify burst-prone areas. The vulnerability of
the excavations has to be considered to identify burst-prone areas
in a mine

2.6.4 Concluding remarks


For the reasons discussed in this section, the CRBSHB’96 had to
be overhauled. Major adjustments are reflected in this reference
book with detailed discussions of excavation vulnerability and
strainburst processes in Volume II. This book provides means for
assessing the excavation vulnerability for all possible rockburst
mechanisms including dynamically loaded falls of ground. It now
differentiates between seismically triggered falls of ground or
triggered strainbursts (when the damage is essentially unrelated to
the source intensity) and dynamically-loaded falls or strainbursts
(when the damage is in part, but not exclusively, related to the
intensity of the remote seismic event).
A burst-prone mine should construct and maintain a mine-wide
rockburst damage database that documents the type of rockburst
damage, damage severity and the rock support system perfor-
mance. This is an important step to assist in selecting rockburst
support based on site-specific experience. Such databases as pre-
sented here are also needed to verify the design approach for a
given mine.

2.7 References

Aki, K., and Richards, P.G. 2002. Quantitative Seismology, 2nd Edition.
Univ. Sci. Books, Sausalito, CA. p.
Alejano, L.R., and Alonso, E. 2005. Considerations of the dilatancy angle in
rocks and rock masses. Int. J. Rock Mech. Min. Sci. 42(4): 481-507.
Araneda, O., and Sougarret, A. 2008. Lessons Learned in Cave Mining at El
Teniente Mine Over the Period 1997-2007. In Proc. Conf. MassMin. pp.
43-52.
Beck, D.A., Reusch, F., Arndt, S., Thin, I., Heap, M., Tyler, B., and Stone, C.
2006. Numerical Modelling of Seismogenic Development During Cave
Initiation, Propagation and Breakthrough. In Deep and High Stress
Mining 2006.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 87
Bewick, R.P. 2013. Fault rupture mechanics under constant normal stress
and stiffness. University of Toronto. Ph.D. Thesis. p.
Bewick, R.P., Kaiser, P.K., and Bawden, W.F. 2014a. DEM simulation of
direct shear: 2. Grain boundary and mineral grain strength component
influence on shear rupture. Rock Mechanics and Rock Engineering
47(5): 1673-1692.
Bewick, R.P., Kaiser, P.K., and Bawden, W.F. 2014b. Shear rupture under
constant normal stiffness boundary conditions. Tectonophysics 634: 76-
90.
Bewick, R.P., Kaiser, P.K., Bawden, W.F., and Bahrani, N. 2014c. DEM
simulation of direct shear: 1. Rupture under constant normal stress
boundary conditions. Rock Mechanics and Rock Engineering 47(5):
1647-1671.
Blake, W., and Hedley, D.G.F. 2003. Rockbursts, case studies from North
American hard-rock mines. SME, Littleton, CO. p. 121.
Board, M., Damjanac, B., and Pierce, M. 2007. Development of a
methodology for analysis of instability in room and pillar mines. In
Deep Mining 07, Proceedings of the Fourth International Seminar on

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Deep and High Stress Mining. Edited by Y. Potvin. pp. 273-282.
Cai, M. 2008. Influence of stress path on tunnel excavation response -
numerical tool selection and modeling strategy. Tunnelling and
Underground Space Technology 23(6): 618-628.
Cai, M., Kaiser, P.K., Cotesta, L., and Dasys, A. 2006. Planning and design
of underground excavations utilizing common earth model and
immersive virtual reality. Chinese Journal of Rock Mechanics and
Engineering 25(6): 1182-1189.
Cai, M., Kaiser, P.K., Morioka, H., Minami, M., Maejima, T., Tasaka, Y.,
and Kurose, H. 2007. FLAC/PFC coupled numerical simulation of AE
in large-scale underground excavations. Int J Rock Mech Min Sci 44(4):
550-564.
Catalan, A., Onederra, I., and Chitombo, G. 2017. Evaluation of intensive
preconditioning in block and panel caving–part II, quantifying the effect
on seismicity and draw rates. Mining Technology: 1-19.
Detournay, E. 1986. Elastoplastic model of a deep tunnel for a rock with
variable dilatancy. Rock Mech Rock Engng 19(2): 99-108.
Diederichs, M.S. 2014. When does Brittle Failure Become Violent? Spalling
and Rockburst Characterization for Deep Tunneling Projects. In
Proceedings of the World Tunnel Congress. pp. 1-10.
Durrheim, J., Roberts, M.K.C., Haile, A.T., Hagan, T.O., Jager, A.J.,
Handley, M.F., Spottiswoode, S.M., and Ortlepp, W.D. 1998. Factors
influencing the severity of rockburst damage in South African gold
mines. J. South Afr. Inst. Min. Metall.: 53-57.
Durrheim, R.J. 2012. Functional specifications for in-stope support based on
seismic and rockburst observations in South African mines. In Deep
Mining 2012. Edited by Y. Potvin, Perth, Australia. pp. 41-55.
Feng, X., Chen, B., Li, S., Zhang, C., Xiao, Y., Feng, G., Zhou, H., Qiu, S.,
Zhao, Z., and Yu, Y. 2012. Studies on the evolution process of
rockbursts in deep tunnels. Journal of Rock Mechanics and
Geotechnical Engineering 4(4): 289-295.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


88 Rockburst phenomena and rockburst damage
Gibowicz, S.J. 2009. Seismicity induced by mining: recent research. In
Advances in Geophysics. pp. 1-53.
Gutenberg, B., and Richter, C.F. 1954. Seismicity of the earth and associated
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

phenomena. Princeton University Press, Princeton. p.


Hanks, T.C., and Kanamori, H. 1979. A moment magnitude scale. Journal of
Geophysical Research: Solid Earth 84(B5): 2348-2350.
Heal, D., Potvin, Y., and Hudyma, M. 2006. Evaluating rockburst damage
potential in underground mining. In Golden Rocks 2006, The 41st U.S.
Symposium on Rock Mechanics (USRMS): "50 Years of Rock
Mechanics - Landmarks and Future Challenges". Paper 1020.
Hedley, D.G.F. 1992. Rockburst handbook for Ontario hardrock mines.
CANMET Special Report SP92-1E. p. 305.
Jenkins, F.M., Conway, G.A., Dwyer, J.G., and Signer, S.P. 2006. 50 years
of rock mechanics research (1955-2005): The effect on safety in U.S.
underground mines. In Golden Rocks 2006, The 41st U.S. Symposium
on Rock Mechanics (USRMS): "50 Years of Rock Mechanics -
Landmarks and Future Challenges", Golden, Colorado. Paper 1177.
Kaiser, P.K., and Cai, M. 2013a. Critical review of design principles for rock
support in burst-prone ground - time to rethink! Keynote. In Ground
Support 2013. Edited by Y. Potvin and B. Brady. pp. 3-38.
Kaiser, P.K., and Cai, M. 2013b. Rockburst damage mechanisms and
support design principles. Keynote. In RaSiM8 Proceedings, Saint-
Petersburg, Moscow, Russia. pp. 349-370.
Kaiser, P.K., and Maloney, S.M. 1997. Scaling laws for the design of rock
support. Pure and Applied Geophysics 150(3-4): 415-434.
Kaiser, P.K., Tannant, D.D., and McCreath, D.R. 1996. Canadian Rockburst
Support Handbook. Geomechanics Research Centre, Laurentian
University, Sudbury, Ontario. p. 314.
Kaiser, P.K., Tannant, D.D., McCreath, D.R., and Jesenak, P. 1992.
Rockburst damage assessment procedure. In Rock Support in Mining
and Underground Construction. Balkema, Rotterdam. pp. 639-647.
Kaiser, P.K., Vasak, P., Suorineni, F., and Thibodeau, D. 2005. New
dimensions in seismic data interpretation with 3-D virtual reality
visualization in burst-prone mines. In RaSiM6 - Sixth International
Symposium on Rockburst and Seismicity in Mines. Edited by Y. Potvin
and M. Hudyma. ACG, Perth, Western Australia. pp. 33-45.
Kaiser, P.K., Yazici, S., and Maloney, S. 2001. Mining-Induced Stress
Change and Consequences of Stress Path on Excavation Stability-A
Case Study. Int J Rock Mech Min Sci 38(2): 167-180.
Karampinos, E., Hadjigeorgiou, J., Hazzard, J., and Turcotte, P. 2015.
Discrete element modelling of the buckling phenomenon in deep hard
rock mines. International Journal of Rock Mechanics and Mining
Sciences 80: 346-356.
Ma, T.H., Tang, C.A., Tang, L.X., Zhang, W.D., and Wang, L. 2015.
Rockburst characteristics and microseismic monitoring of deep-buried
tunnels for Jinping II Hydropower Station. Tunnelling and Underground
Space Technology 49: 345-368.
Manouchehrian, A., and Cai, M. 2016. Simulation of unstable rock failure
under unloading conditions. Canadian Geotechnical Journal 53(1): 22-
34.
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 89
McGarr, A. 1984. Scaling of ground motion parameters, state of stress, and
focal depth. J. Geophysical Research 89: 6969-6979.
Mendecki, A.J. 2013. Characteristics of seismic hazard in mines. In The 8th
Rockburst and Seismicity in Mines Symposium, Russia.
Mendecki, A.J. 2016. Mine Seismology Reference Book, Seismic Hazard.
Institute of Mine Seismology, Australia. p. 88.
Morissette, P., Hadjigeorgiou, J., Thibodeau, D., and Potvin, Y. 2012.
Validating a support performance database based on passive monitoring
data. In Proceedings of the Sixth International Seminar on Deep and
High Stress Mining. pp. 41-55.
Ortlepp, W.D. (ed). 1997. Rock Fracture and Rockbursts – An Illustrative
Study. The South African Institute of Mining and Metallurgy,
Johannesburg.
Ortlepp, W.D. 2000. Observation of mining-induced faults in an intact rock
mass at depth. Int. J. Rock Mech. Min. Sci. 37(1-2): 423-436.
Ortlepp, W.D., and Stacey, T.R. 1994. Rockburst mechanisms in tunnels and
shafts. Tunnelling and Underground Space Technology 9(1): 59-65.
Potvin, Y. 2009. Strategies and tactics to control seismic risks in mines. J.
South Afr. Inst. Min. Metall. 109(March): 177-186.

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Robbins, R.J. 2010. Dealing with rock mechanics challenges in a machine
bored tunnel. In Proc. 44th US Rock Mechanics Symposium and 5th
U.S.-Canada Rock Mechanics Symposium, Salt Lake City, UT.
Rojat, F., Labiouse, V., Descoeudres, F., and Kaiser, P.K. 2001. Brittle rock
failure at the Loetschberg. Civil engineering department of EPFL,
Lausanne. p. 31.
Salamon, M.D.G. 1970. Stability, instability, and design of pillar workings.
Int. J. Rock Mech. Min. Sci. 7: 613-631.
Salamon, M.D.G. 1983. Rockburst hazard and the fight for its alleviation in
South African gold mines. In Rockbursts: prediction and control. The
Institution of Mining and Metallurgy. pp. 11-36.
Salamon, M.D.G. 1999. Strength of coal pillars from back-calculation. In
Proceedings of 37th US rock mechanics symposium, Vail. 1 pp. 29-36.
Stacey, T.R. 2016. Addressing the Consequences of Dynamic Rock Failure
in Underground Excavations. Rock Mechanics and Rock Engineering
49(10): 4091-4101.
Suorineni, F.T., Kaiser, P.K., Mgumbwa, J.J., and Thibodeau, D. 2011.
Mining of orebodies under shear loading Part 1 – case histories. Mining
Technology 120(3): 137-147.
Suorineni, F.T., Mgumbwa, J.J., Kaiser, P.K., and Thibodeau, D. 2014.
Mining of orebodies under shear loading Part 2 – failure modes and
mechanisms. Mining Technology 123(4): 240-249.
Tarasov, B., and Potvin, Y. 2013. Universal criteria for rock brittleness
estimation under triaxial compression. International Journal of Rock
Mechanics and Mining Sciences 59(4): 57-69.
Vallejos, J., and McKinnon, S. 2008. Guidelines for Development of Re-
entry Protocols in Seismically Active Mines. In 42th US Symp. Rock
Mech, San Francisco. Paper 08-097.
Wang, X., and Cai, M. 2015. Influence of wavelength-to-excavation span
ratio on ground motion around deep underground excavations.
Tunnelling and Underground Space Technology 49: 438-453.
Draft manuscript – Copyright protected – Cai and Kaiser 2018
90 Rockburst phenomena and rockburst damage
Wang, X., and Cai, M. 2016. FLAC/SPECFEM2D coupled numerical
simulation of wavefields near excavation boundaries in underground
mines. Computers & Geosciences 96: 147-158.
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Zhang, C., Feng, X.-T., Zhou, H., Qiu, S., and Wu, W. 2012. Case histories
of four extremely intense rockbursts in deep tunnels. Rock mechanics
and rock engineering 45(3): 275-288.
Zhao, X.G., and Cai, M. 2010. A mobilized dilation angle model for rocks.
Int. J. Rock Mech. Min. Sci. 47(3): 368-384.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Chapter Three
Design principles and methodology
Synopsis
A detailed description of rock support functions in burst-prone
grounds is introduced after a general discussion of design princi-
ples in geotechnical engineering. Emphasis is placed on the need
for proper connections between support components to create an
integrated support system that reinforces, retains, and holds pack-
ages of stress-fractured rock (called gabions) in place.
In the context of rock support selection for burst-prone ground,
seven guiding principles are presented for ground control deci-

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
sion-making and rock support selection. Standard support design
methodology and their limitations are briefly reviewed.
An overview of the rockburst support design process, including
data collection, selection of design domains and the evaluation of
support demands and capacities is provided. Support capacity
consumption and restoration are newly introduced as important
design criteria.
Finally, the mitigation of rockburst damage caused by excavation
failure and dynamic disturbances is briefly discussed and design
concepts to mitigate rockburst damage caused by a self-initiated
rockburst (strainburst), a seismically triggered strainburst, and
dynamically loaded rockbursts are introduced.

3 Design principles and methodology............................................. 93


3.1 Engineering principles ......................................................... 93
3.1.1 Identification of relevant failure mechanisms................ 93
3.1.2 Identification of causes and severity of failure .............. 93
3.1.3 Identification of a ‘design event’ .................................. 94
3.1.4 Determination of a safety margin .................................. 95
3.1.5 Role of condition monitoring and design for multiple lines
of defense ................................................................................... 95
3.2 Support and its function in stress-fractured ground ............... 96

Draft manuscript – Copyright protected – Cai and Kaiser 2018


92 Design principles and methodology

3.2.1 How is a support loaded in bursting ground? ................. 96


3.2.2 Rock support functions ................................................. 97
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

3.2.3 Integration of support functions to form reliable support


systems 101
3.3 Rockburst support design principles.................................... 103
3.3.1 Avoid rockbursts......................................................... 104
3.3.2 Use deformable support components ........................... 105
3.3.3 Address the weakest link............................................. 105
3.3.4 Create an effective, deformable and integrated support
system 106
3.3.5 Principle of simplicity ................................................. 107
3.3.6 Efficient support systems ............................................ 107
3.3.7 Observational approach to ‘anticipate and adapt’......... 108
3.4 Support design methodology .............................................. 109
3.4.1 Analytical support design............................................ 109
3.4.2 Empirical rock support design ..................................... 110
3.4.3 Numerical rock support modelling .............................. 110
3.4.4 Observational design................................................... 112
3.5 Overview of rockburst support design process .................... 113
3.5.1 Data collection ............................................................ 113
3.5.2 Zoning – design domains ............................................ 114
3.5.3 Evaluation of support demands ................................... 114
3.5.4 Evaluation of rock support system capacity ................. 115
3.5.5 Support capacity consumption and restoration............. 115
3.5.6 Design verification and modification........................... 117
3.6 Mitigation of rockburst damage caused by excavation failure
and dynamic disturbances .............................................................. 118
3.7 References .......................................................................... 119

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 93

3 Design principles and methodology

3.1 Engineering principles


Geotechnical or geomechanics engineering entails several funda-
mental principles that are summarized next. For a proper design, it
is necessary to:
- recognize or identify the relevant or physically possible failure
mechanisms that need to be analysed;
- identify the cause for and severity of the failure process to establish
possible driving forces or deformation demands;
- identify and define a ‘design event’, i.e., the intensity that needs to
be ‘survived’ or is acceptable (allowed) for a given design period;

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
and
- select a safety margin for each possible failure mechanism.

3.1.1 Identification of relevant failure mechanisms


For geotechnical engineering designs, it is always necessary to
identify possible and likely rock failure mechanisms. By analogy
to pit slope stability assessment, for example, one needs to antici-
pate the likely failure modes and ask: is a slope failing by toppling,
wedge failure or sliding on a planar or circular surface? For rock-
burst damage, the mode of stress-fracturing needs to be estab-
lished; e.g., is the failure caused by spalling, buckling, yielding, or
shearing? This aspect is introduced in the previous chapter and
discussed in more detail in Chapter 1 of Volume II.

3.1.2 Identification of causes and severity of failure


What are the failure causing drivers? In many engineering disci-
plines, engineering standards define design approaches and as-
sumptions, e.g., in civil engineering, regional earthquake loads
(accelerations), hundred-year snow or wind loads, flood levels,
etc., are specified. By analogy, for burst support design, as no
formal engineering standards exist, the mining-induced, static
stress loading conditions have to be defined at the point of failure,
i.e., the stress level and the vulnerability of an excavation, and the
dynamic disturbances in the form of dynamic stress changes and
ground motions. The aspect of excavation vulnerability is covered
in detail in Chapter 1 of Volume II.
Draft manuscript – Copyright protected – Cai and Kaiser 2018
94 Design principles and methodology

For design purposes, it is not necessary or advisable to establish


the actual prevailing conditions at a given location and time but to
establish the equivalent to a x-year loading period, where x is the
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

return period that is considered to be appropriate for a given exca-


vation (e.g., x = 1 to 2 years for temporary drifts such as drill
horizons in caving operations, x = 10 to 20 years for infrastructure,
and x = 50 to 100 years for main mine accesses such as ramps and
shafts). Furthermore, for a design in bursting ground it is neces-
sary to differentiate between factors causing failure and factors
affecting the severity of failure. The former includes factors af-
fecting the rockburst potential (RBP) and the latter includes fac-
tors affecting the rockburst severity (RBS).

3.1.3 Identification of a ‘design event’


The first step in a design is to establish meaningful ‘design events’.
The term ‘design event’ has the same meaning as, for example, the
100-year flood level ‘design event’ for a water retention structure.
‘Design events’ are conditions for an engineered structure to
survive. They are defined in terms of a single factor or multiple
factors that are relevant for a design. For example, for support
against damage from a remote seismic event, the intensity (magni-
tude) of the seismic event, the volume rock containing this event
(or the location relative to the excavation that is to be supported),
and the anticipated radiation pattern (i.e., the ground motions
(PGV and PGA or Dsd )) are of significance.
A ‘design event’ defines what the extreme situations are, condi-
tions that need to be survived, and other design conditions or
assumptions have to be specified to arrive at a sound engineering
design. Amongst other factors, it has to be defined:
- what the vulnerability of an (supported) excavation is, i.e., the
safety margin (or factor of safety) before the excavation is af-
fected by a dynamic disturbance;
o for strainbursting ground, it is necessary to define the
eventual prevailing stress and mine stiffness condition as
well as the anticipated depth of failure and rock mass
(post-peak) behaviour;
o for shakedown failure, the anticipated geometry (or mass)
of a potential shakedown volume is to be defined.
Therefore, a ‘design event’ can rarely be defined in terms of the
maximum seismic event magnitude alone. The excavation status,

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 95

the loading system stiffness, and the effectiveness of the installed


support have to be described. This book provides an approach to
establish design controlling drivers of failure. Aspects dominating
rockburst damage caused by dynamic excavation failures are
discussed in Volume II and those dominated by dynamic disturb-
ances from remote seismic design events are covered in Volume
III.
There are no formal engineering standards to define ‘design
events’ for burst-prone ground. However, a sound engineering
design has to start with a list of design assumptions that include a
description of what a selected support system needs to survive.
This book provides guidance on how to establish relevant assump-
tions (‘design events’) that can be used to specify meaningful
design conditions.

3.1.4 Determination of a safety margin

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
As in all geotechnical engineering practises, the safety margin or
risk of failure is assessed and measured by either a factor of safety
(FS), a safety margin (SM) or a probability of failure (Pf). This
book primarily uses the FS or the SM approach, whereby the
demand from a design event is compared with the capacity of the
mitigating ground control measure – the rock support system.

3.1.5 Role of condition monitoring and design for multiple lines


of defense
Uncertainties from the variability of natural materials, the com-
plexity of failure processes and the need for flexibility in mining
would require uneconomical factors of safety. Hence, it is a com-
mon practice in geomechanics to manage uncertainty by two
means: (a) reduce the risk by narrowing the variability in demand
and capacity through performance monitoring, and (b) reduce the
risk by implementing several lines of defence, i.e., by selecting a
support system that will not fail when one component of the sys-
tem is overloaded.
The implications of this engineering approach are that support
needs to be designed for various possible conditions (static and
dynamic) and for various possible failure mechanisms (e.g., gravi-
ty-driven wedge failure, stress-driven rock fracturing, seismic
shaking, ejection, etc., occurring sequentially or simultaneously).
Furthermore, as indicated above, the design has to take into ac-

Draft manuscript – Copyright protected – Cai and Kaiser 2018


96 Design principles and methodology

count possible demands or ‘design events (loads)’ that are not


equal to those actually measured. Possible demands are deter-
mined by establishing a reasonable representation of future reali-
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

ties. As the slip direction and thus the radiation pattern of the
‘seismic design event’ is unknown (for more detail see Volume
III), ground motions, for example, may be estimated by scaling
laws, which imply equal and maximum wave radiation in all
directions. This book, as the 1996 handbook (Kaiser et al. 1996),
utilizes scaling laws or ground motion equations to determine
anticipated ‘design’ ground motions from large remote seismic
events, i.e., to determine dynamic disturbances in the form of PGV,
PGA and dynamic stress pulses Dsd.
The use of scaling laws, if executed properly, forms a sound basis
for a support selection but not for a forensic analysis and an exca-
vation damage assessment (e.g., not for correlating PGV from a
scaling law to an observed rockburst damage; see Volume III). For
forensic analyses, the excavation vulnerability and the actual
ground motion need to be predicted (e.g., by synthetic ground
motion models) and then compared with the observed damage. A
synthetic ground motion assessment approach is introduced in
Volume III of this book to assist designers in an excavation and
support damage assessment.

3.2 Support and its function in stress-fractured


ground
3.2.1 How is a support loaded in bursting ground?
When hard brittle rock fails by stress-fracturing, the rock mass
bulks and deforms into the excavation. This bulking deformation
loads support components in two ways (Figure 3-1): (top) it trans-
fers loads (force F) directly or indirectly via the retaining compo-
nents (e.g., mesh or shotcrete) to the plate of the holding elements;
and (bottom) it strain rockbolts inside the rock mass by a relative
push motion ‘from behind’, i.e., by relative block movements
during rock mass bulking as schematically illustrated by Figure
3-1. These differential displacements typically impose extension
and some shear strain on the rock reinforcement.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 97

Ly = 100 to 150 mm
F
• Pull test: displacement at plate
F rebar MCB
F=0
d

• Rock mass bulking:


loads bolts in a push motion

Figure 3-1 Rockbolt loading process: (top) plate loading force F via deformation of a
retain system; and (bottom) by rock mass loading causing differential block movement

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
and rockbolt straining (Kaiser 2017).
The first loading process is reproduced by pull-out tests and the
respective results reveal the capacity of a bolt to resist plate load-
ing during a rock burst. The top diagrams show that a rebar is
mostly strained near the plate, whereas modified cone bolts (MCB)
are deforming primarily by cone ploughing toward the end of the
bolt.
The second loading process cannot be easily tested in the field.
This process is typically established in the laboratory by indirect
loading tests, where the performance of a bolt depends on the
relative movement patterns between individual rock blocks. If a
bolt is uniformly strained, it will attain its maximum capacity but
if a localization of movement occurs, its capacity will be severely
reduced. Hence, for an optimal performance of a support system,
it is necessary to control the bulking process and prevent dis-
placement localizations.

3.2.2 Rock support functions


From an operational perspective, a rock support is installed to
ensure personnel safety, to safeguard mining equipment, to keep
the excavation functionally open, and to ensure continuous mine
production. Similar to understanding excavation failure processes,
it is necessary to understand and differentiate between support
functions in order to arrive at robust and effective support systems.
Not every support component, i.e., bolt, mesh or shotcrete, serves
the same rock support function.
Draft manuscript – Copyright protected – Cai and Kaiser 2018
98 Design principles and methodology
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Figure 3-2 Three functions of a support systems: reinforce, retain, and hold (Kaiser et al.
1996).
The mechanics of rock support is complex and most models trying
to simulate the rock/support interaction mechanisms are flawed as
they do not fully capture the load-sharing interaction among sup-
port components of a rock support system and the interaction
between rock and rock support. Kaiser et al. (2000) defined three
primary support functions as illustrated by Figure 3-2: (1) rein-
force the stress-fractured or yielded rock mass to strengthen it and
to control bulking, (2) retain broken rock to prevent unravelling
between tendons, and (3) hold the broken rock in the retaining
element and securely tie it back to stable1 ground. In addition to
these three support functions, two very important conditions have
to be met: (1) all support elements, providing the three functions,
must always work together, i.e., they have to be able to share loads
while the support is being deformed, and (2) they have to be well
connected such that the connections do not fail by creating a weak
link in the support system. Holding and retaining components
must be well connected to ensure system integrity and stability,
e.g., by using straps, large plates, strong threads, etc. This forth
function has been added to the cartoon that is used throughout this
book (Figure 3-3).

1
Stable ground does not mean elastic ground; anchorage is possible in a yielded rock
mass as long as its load-bearing capacity has not been exhausted by excessive loss of the
cohesive component of rock mass strength.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 99

Yielding bolt/hold
Arching
Rock mass strength enhanced
Fractured rocks by support confinement

Enhanced post-peak
strength of stress-
Mesh-reinforced fractured rock
shotcrete/retain

Strap/connect
Rebar/reinforce

Figure 3-3 An illustration of an integrated rockburst support system that possesses the

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
required support functions: reinforce, retain, hold, and connect.

Reinforce
The goal of reinforcing the rock mass is not only to strengthen it,
thus enabling the rock mass to support itself (Hoek and Brown
1980), but also to control the bulking process as rockbolts/rebar
prevent fractures from opening and propagating by extension.
Rock reinforcement is achieved by installing rebar, mechanical
rockbolts, or cablebolts in a grid pattern. As long as they do not
yield excessively, fully grouted rebar and cablebolts are most
effective in controlling bulking as they increase the shear re-
sistance of joints and enhance the interlock of strong rock blocks.
A rock mass reinforcement can raise the trigger limit for rockburst
damage and help maintain a high post-peak shear resistance in the
fractured rock (Figure 3-3). Fully grouted reinforcement compo-
nents are stiff and can fail when strained beyond their peak
strength. It is, however, inappropriate to conclude that fully grout-
ed rebar are not suitable as rock support components in burst-
resistant support systems. The bulking control role of a rebar is
maintained between breaks and this prevents widespread bulking.
This helps to reduce convergence due to bulking and protects
other bolts from excessive straining. Of course, a broken rebar
alone cannot hold the reinforced rock in place. Other support
components, which have the ability to retain and hold, are needed
to create a functional integrated rock support system.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


100 Design principles and methodology

Retain
Rock fracturing is inevitable under high static or dynamic stress
loading and the resulting rock fragments tend to unravel between
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

the tendons if they are not properly retained. An important func-


tion of a support system therefore is to retain the fractured rock
near the excavation boundary and to transfer its weight or load to
the holding elements. Because confined fractured rock mobilizes
more frictional resistance, the retention component also helps to
strengthen the broken rock mass and to provide confinement to
deeper rock.
Widely used retaining elements are wire mesh and strap, plain and
fibre/mesh-reinforced shotcrete, embedded steel arches or girders,
or cast-in place concrete. In burst conditions, shotcrete needs to be
reinforced by adding fibres and preferably mesh to increase its
tensile strength and toughness. While fibres are effective at small
tensile strains, mesh-reinforced shotcrete or mesh over shotcrete is
required when tensile cracking is localized (e.g., at pillar noses) or
point loading due to rock impact during dynamic loading is ex-
pected.
Hold
A well-retained and reinforced rock arch may excessively deform
or may be at risk of failure by shearing through the reinforced rock.
Hence, holding elements are needed to tie the retaining elements
and the reinforced broken rock to stable ground. It has to yield
when deformed during rock mass bulking and may have to dissi-
pate the dynamic energy. Strong holding elements (cables) are
typically used to prevent the triggering of gravity-dominated falls
of ground. These holding elements with high load and deformation
capacities (and thus high energy dissipation capacities) also assist
in surviving dynamic loads due to acceleration from seismic shak-
ing. When rockburst damage is anticipated, yielding holding ele-
ments such as deboned cablebolts, conebolts, or any other proven
yielding bolt (e.g., D-bolts) may have to be added to the support
system.
Connections
In an effective support system, the three support functions act
together by transferring and sharing loads and by deforming to-
gether to simultaneously provide confinement and dissipate ener-
gy. The full load and deformation capacities or the energy dissipa-
tion capacity of a support system is only achieved if all rock
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 101

support elements are well connected, forming a fully integrated


rock support system. Unfortunately, connections between retain-
ing and holding elements frequently constitute the weakest link in
a support system. These connections are often the cause for a
support system’s failure, and thus deserve special attention in
support details. The reader is referred to Chapter 4 in Kaiser et al.
(1996) for more information on connections.

3.2.3 Integration of support functions to form reliable support


systems
The functionality of rock support components is interlinked and
these components depend on each other to ensure their full capaci-
ty. Figure 3-4 illustrates that all four support functions (reinforce,
retain, hold, and connect) are needed in an effective rockburst
support system no matter what the damage mechanism or the

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
damage severity might be. However, the relative contributions of
each component to a system’s performance will differ and, as will
be discussed in Chapter 4, will change as a support system is
being deformed. As a matter of fact, some of the support capacity
is consumed while the support is deformed and the ‘remnant’
capacity decreases with increasing deformation. Furthermore, the
capacity of a support system depends on the installation sequence
of the various components making up the system because each
component has a different displacement tolerance.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


102 Design principles and methodology
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Figure 3-4 Rockburst damage mechanism, damage severity, and required support func-
tions.
The elements of an integrated rock support system for burst-prone
ground were introduced earlier and are illustrated by Figure 3-3.
This support system retains and holds fractured rocks in place as a
gabion retains boulders in a steel mesh to form a retaining wall
(Figure 3-5a). In the damaged rock mass in the immediate vicinity
of an excavation (wall or back), the retaining elements combined
with relatively short rebar form a ‘gabion’ of stress-fractured rock
as the cartoon in Figure 3-5b illustrates. These gabions form a
support arch by mobilizing the strength of the reinforced fractured
rock. This support arch then provides confinement and enhances
the rock mass strength outside the fractured zone. It may be neces-
sary to tie the ‘gabions’ with ‘holding’ elements back to stable
ground to ensure stability of the overall ‘gabion’ arch. This is
schematically illustrated by the red bolts or cables in Figure 3-5b.
Proper support of the individual ‘gabions’ containing broken rock
is a prerequisite for the formation of an effective ground arch.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 103

(a) (b)
Figure 3-5 (a) Gabions used to stabilize a steep part of a rock slope, and (b) deformable
support system for ground control in wall of adrift indicating resistance forces of the
reinforced fractured rock to resist tangential straining (vertical arrows) and to provide
confinement to the surrounding rock mass (horizontal arrows) (Kaiser 2017).
In summary, rock support systems are used in burst-prone grounds

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
to withstand loads, to minimize deformations by reinforcement, to
accommodate large displacements and dissipate dynamic energy
by yielding. Therefore, rockburst resistant support systems must
be designed to provide sufficient load, displacement, and energy
capacities to meet the corresponding demands. The two features of
capacity and demand are covered separately in this book even
though they are interlinked. The capacity of support components is
covered in Chapter 4 of this volume and load, displacement, and
energy demands are covered in Volumes II and III.

3.3 Rockburst support design principles


In underground construction, strategy is the art of commanding the entire
mining or tunneling operation. Tactics, on the other hand, are the
actions aiming at the achievement of a goal by using various tools
for the construction and during the life of an excavation. Engineers
are frequently forced to find tactical solutions as strategic thinking de-
mands long-term planning. It is often too late to get out of a reactive
mode when dealing with rockbursts.
Ralph Waldo Emerson2 stated “As to methods there may be a
million and then some, but principles are few. The man who
grasps principles can successfully select his own methods. The

2
For quotes by Ralph Waldo Emerson visit:
https://www.goodreads.com/author/quotes/12080.Ralph_Waldo_Emerson

Draft manuscript – Copyright protected – Cai and Kaiser 2018


104 Design principles and methodology

man, who tries methods, ignoring principles, is sure to have trou-


ble.” Realizing the importance of understanding rockburst support
design principles, Cai and Champaigne (2009) and Cai (2013)
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

grouped practical experiences into seven principles (Figure 3-6) to


mitigate rockburst and to select a support. By understanding these
principles, the ability to safeguard workers and company property
can be increased. These core principles must guide support design.

1. Avoid
rockbursts

7. Anticipate 2. Use de-


and be adapt- formable
able support

6. Use effi- 3. Address the


cient systems weakest link

4. Create effective
5. Simplicity integrated
support system

Figure 3-6 Summary of seven rockburst mitigation and support selection principles
(modified after Cai (2013)).

3.3.1 Avoid rockbursts


The supreme goal in rockburst management is to avoid rockburst
conditions, thereby eliminating or at least minimizing the need for
burst-resistant support. The best strategy is to stabilize the rock
without fighting loads and stresses by the use of a heavy rock
support. Ržiha (1874), a famous nineteenth century tunnelling
engineer, once commented “The true art in tunnelling lies in the
anticipation of the development of large rock pressure, which is
far more effective than to find the means of resisting rock pres-
sures which have already developed.” Sun Tzu (1963), a military
strategist in ancient China, said, “To fight and conquer in all your
battles is not supreme excellence; the supreme excellence consists
in breaking the enemy's resistance without fighting.”

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 105

Methods to avoid rockburst risks amongst others include changing


drift shape and locations, changing stope size/shape and mining
sequencing, minimizing extraction ratios, avoiding stress raisers,
and potentially switching to a different mining method.

3.3.2 Use deformable support components


When brittle rock fails, it causes large rock bulking deformations
and if the failure process is unstable, energy is released. In addi-
tion, the rock may be subjected to stress pulses, ground motions,
and energy transfer from a remote seismic event. The installed
rock support components must therefore be able to reinforce the
rock so as to minimize rock mass bulking. At the same time, they
must be deformable to comply with bulking deformations and able
to absorb dynamic energy; they have to accommodate static and
dynamic rock deformations. It is often more economical to in-

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
crease the deformation rather than the load capacity of a support
element to control rockburst damage. Deformable support compo-
nents are able to tolerate large tunnel convergence without ‘self-
destruction’ while absorbing dynamic energy (product of load
capacity and deformability). A yielding rock support component
must be in harmony with the damaged rock surrounding the exca-
vation.

3.3.3 Address the weakest link


A chain is only as strong as its weakest link. In conventional rock
support systems, the retaining element is often the weakest link
because the retention systems or the connections to the tendons
fail; e.g., if weak mesh is used, sharp-edged steel plates cut the
mesh, the mesh overlaps are insufficient, bolts fail at the threaded
section, plates fail prematurely, and the connections are lost.
Many opportunities exist to enhance the support system capacity
by addressing the weakest link issue. Unfortunately, rock support
design and quality assurance procedures in underground construc-
tion mostly focus on the load and energy capacity of individual
rockbolts. The failure of rock mass between the tendons and the
negative impact of this failure on the capacity of individual bolts
and the overall rock support system is often neglected.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


106 Design principles and methodology

3.3.4 Create an effective, deformable and integrated support


system
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

An effective rock support system is comprised of rock support


components that provide all support functions (reinforce, retain,
hold, and connect) required to control burst-prone ground. Some
components act in series (e.g., transferring load from the retention
system to the bolts) or in parallel (e.g., by simultaneously deform-
ing different bolt types). Some of these components fulfill multi-
ple functions but they may be strong in one aspect and weak in
others. Various support elements must be combined to form a
deformable, integrated support system to maintain optimal support
functionality as the support is deformed. The connections have to
be strong even after large wall convergence, bolts with limited
displacement capacity have to be able to shed load, and holding
elements must be able to assume and maintain this load at relative-
ly large deformations. The overriding principle for the creation of
an integrated support system is to maintain deformation compati-
bility. If deformation compatibility is maintained, a support sys-
tem composed of a given set of support components will attain the
highest possible energy dissipation capacity.
A fundamental requirement to ensure deformation compatibility is
to minimize displacements or strains and to prevent localization of
deformations. For example, holding elements (such as yielding
bolts and debonded cables that can accommodate deformation)
should be combined with reinforcing elements (such as rebar to
control bulking) and competent surface support elements (such as
mesh and shotcrete to provide confinement). This often leads to
some apparently inconsistent requirements as some components
act in parallel (e.g., bolts), whereas others act in series (e.g., mesh
or shotcrete and bolts). For example, a shotcrete liner (or arch)
may be stiff and fail at relatively small displacements compared
with the displacement capacity of bolts. As a consequence, the
maximum potential retention capacity of shotcrete may be lost
long before the ultimate energy dissipation capacity of the bolts is
reached.
With respect to shotcrete, for example, shotcrete arches with
superior support pressure capacity might be beneficial to prevent
rockburst damage (i.e., to decrease the vulnerability of an excava-
tion to rockburst damage or the rockburst potential). On the other
hand, if rockburst damage cannot be prevented, shotcrete panels

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 107

rather than arches should be used to ensure deformation compati-


bility between the tendons and the retention system. These com-
peting goals lead to solutions with longitudinal slots in shotcrete
as frequently used in civil tunnels. Such slots, when open, provide
deformable panels and, when closed, provide radial support pres-
sure.

3.3.5 Principle of simplicity


Simplicity is powerful. Rock support elements should be relatively
easy to manufacture and simple to install. Regardless of how
effective they are, if a rock support element is complicated to
manufacture and the cost is high, operators will be reluctant to use
them. If they are difficult to install and production is adversely
affected, acceptance and installed support quality will suffer. If the
quality cannot be ensured, the support will not work as it is in-

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
tended and designed.
When it comes to rock support in burst-prone ground, it is always
beneficial to “Make everything as simple as possible, but not
simpler”, a statement attributed by R. Sessions (an American
composer) to A. Einstein in 1950. ‘No simpler’, in terms of rock
support systems for burst-prone ground, unfortunately means a
rather complex mix of compatible support components to form a
support system that can be installed efficiently, upgraded easily,
and maintained as mining-induced deformations, including defor-
mations from rockbursts, consume part of the installed support
capacity.

3.3.6 Efficient support systems


In addition to being effective, a support system must be efficient,
i.e., it has to be cost-effective and rapid to install during regular
mine operations. Splitting the support cycle to render the support
system more cost-effective can frequently accommodate the latter.
With respect to the former, there still exists a misconception in
that rockburst-resistant support is considered to be expensive.
Whereas mining companies aim at minimizing costs, they do not
wish to compromise safety and hamper productivity by unneces-
sary rehabilitation and production delays. The consequence of
rockbursts can be extreme, ranging from damage to underground
openings requiring rehabilitation, damage to mining equipment,
loss of production, permanent loss of ore, and may even lead to
injury and fatalities. While support may not prevent a rockburst, it
Draft manuscript – Copyright protected – Cai and Kaiser 2018
108 Design principles and methodology

can lower the consequences and thus the impact of an undesirable


event. The costs associated with a rockburst damage can be ex-
tremely high. For example, it is estimated that the rehabilitation
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

cost is 10 to 20-times higher than the initial support cost in under-


ground hard rock mines. A major rockburst may shut down the
mine production or a tunnel advance for an extended period of
time. Furthermore, if an area has been affected by a seismic im-
pact but survived, the safety margin may have been lowered lead-
ing to increased risks in the future or a demanding preventive
support maintenance may be required. In other words, even if the
price tag for rockburst damage is high, the cost of preventing it by
using a rockburst resistant rock support in the first place can make
this remarkably attractive. Experiences from numerous damage
cases show that prevention and control in burst-prone ground
often offers the most cost-effective solution.

3.3.7 Observational approach to ‘anticipate and adapt’


The last principle presented in Figure 3-6 advocates that it is
essential for risk management in burst-prone ground to ‘anticipate
and adapt’. A high price is to be paid if the unexpected is not
anticipated and operations do not adapt to lessons learned.
The rockburst potential and severity change spatially and tempo-
rarily and it is unrealistic to have a single design that does not
need modification over the mine’s life. Hence, the support system
must be responsive to a variety of ground conditions and changing
demands, in particular to the accumulation of displacements from
the time of a support installation. The art of rock support in burst-
prone ground is not to rely on the low likelihood of unexpected
ground behaviours, but on the readiness to manage the potential
consequences with an effective, robust rock support system. The
adaptability principle demands that observations are continuously
used to verify and to improve the adopted support system. In this
book, we introduce the concept of proactive support maintenance
(PSM) for this purpose. Once a support system has been installed,
PSM utilizes monitoring information (such as cumulative co-
seismic straining, cumulative convergence due to mining and
repeated dynamic loading (‘hammering’)) as indicators of support
consumption to establish when the remnant support capacity is
insufficient and support upgrading by proactive support mainte-
nance is required.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 109

3.4 Support design methodology


Unlike in other engineering disciplines, such as steel and concrete
structure design, there are no formal design standards or codes that
can or must be followed when designing rock support systems for
burst-prone ground. Commonly adopted rock support design
methods, ranging from analytical, empirical, numerical, to obser-
vational methods, are briefly discussed below and summarized in
Figure 3-7. Whereas they might be adapted, they often do not lead
to sound engineering designs for support in burst-prone ground.

Rock support design


methods
Observational

Empirical Analytical/Rational Numerical

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Terzaghi's Closed form
FEM
rock load solution

Depth of
RMR FDM
failure

Wedge
Q-system DEM
analysis

US Corps of
Key block DDA
Engineers

Open stope BEM

FEM/DEM
combined

Figure 3-7 Summary of commonly used rock support design methods.

3.4.1 Analytical support design


For stress-controlled instability problems, analytical design meth-
ods are often based on the closed-form solution for circular exca-
vations in a homogeneous and isotropic ground and in hydrostatic
stress fields. Furthermore, the rock mass is assumed to behave in
an elasto-plastic or strain-softening elasto-brittle plastic manner
with failure envelopes governed by either the Mohr-Coulomb or
the Hoek-Brown failure criterion (Hoek and Brown 1997). These
methods have not been adapted for dynamic conditions.
Structural-controlled instabilities, such as wedges falling from
roof or sliding out of sidewalls formed by intersecting joints or
Draft manuscript – Copyright protected – Cai and Kaiser 2018
110 Design principles and methodology

bedding planes, are identified and analyzed by means of the stere-


ographic projection method. The support in the form of reinforce-
ment can be readily selected by statics and dynamic analyses
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

assuming rigid rock blocks. For the latter, the gravitational accel-
eration is enhanced for dynamic shaking (see Volume III). The
rockbolts or anchors used for roof wedges or roof beams should
provide a sufficient load capacity to support the weight of the
failing ground volume under static load and dynamic acceleration.
Tools such as UNWEDGE (RocScience Inc.) and the key-block
analysis program (Shi 1992) can be used to carry out the design
analysis with enhanced gravitational forces.

3.4.2 Empirical rock support design


Most empirical support design methods are based on widely
adopted rock mass classification systems, such as Terzaghi's rock
load guidelines (Terzaghi 1946), the RMR- (Bieniawski 1984) or
the Q-system method (Barton et al. 1977), and the US Corps of
Engineers (1980)' rules of experience. Specific empirical methods
were developed for support design of large-scale underground
caverns (Cording et al. 1971) and mine stopes (Mathews et al.
1981; Potvin 1988).
When using empirical methods, one must consider their limita-
tions (Palmstrøm and Broch 2006). As empirical methods were
developed based on data from project-specific data, they may lead
to inadequate ground support if the underlying database is not
representative. Furthermore, most empirical methods do not fully
respect the complex stress paths experienced by excavations in
mines. As far as the authors are aware, no empirical support de-
sign method has been tailored for the design of a support in burst-
prone ground in underground mines. Even the Q-system, which
uses the stress reduction factor (SRF) to anticipate rockbursting, is
highly restrictive. It is only applicable to anticipate strainbursts
during tunnel advance and account for (strain-) bursting in rela-
tively stiff loading conditions (single or multiple tunnel advances
without mining-induced stress changes).

3.4.3 Numerical rock support modelling


If more sophisticated designs are justified for technical or eco-
nomic reasons, numerical design methods may be helpful but they
often are not suitable for support design in burst-prone ground;
e.g., the GSI (Geological Strength Index) system (Hoek et al. 1995)
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 111

was developed for slope and tunnel design hence it cannot be


applied without modification to burst-prone ground.
Numerical methods, such as Finite Element Method (FEM),
Boundary Element Method (BEM), Finite Difference Method
(FDM), Discontinuous Deformation Analysis (DDA), and Distinct
Element Method (DEM), have been widely used in the design of
underground excavations. When using numerical tools, complex
geometries and ground conditions such as geology, in-situ, and
mining-induced stresses, as well as nonlinear and non-elastic
material behaviour can be modelled. These tools also offer means
to assess loading and straining of support components (rockbolts,
anchors, shotcrete, or concrete linings, etc.) and to investigate the
impact of a support on the rock mass behaviour. However, where-
as the extent of excavation-induced yielding or fracturing, the
elastic and plastic deformations imposed on the support, and the

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
loading of specifying rock support components can be calculated,
the straining and displacements from geometric bulking of stress-
fractured ground cannot be properly simulated by continuum
models. For brittle failing, bulking rock masses, the most severe
deficiency of continuum models lies in the fact that the defor-
mations imposed on the support elements due to rock bulking are
generally underestimated and dilation parameters cannot be used
to adjust for this deficiency. Hence, numerical approaches often
tend to underestimate support loads.
In burst-prone ground, the field and mining-induced stresses are
high and when coupled with seismic-wave-induced stresses, nu-
merical design analyses may help to assess the impact on the
extent of rock mass failure. However, the implicit result of the
adopted continuum approach does not account for observed be-
haviour whereby stress-fracturing during rockbursts may occur in
relatively large increments rather than by gradual spalling. This
leads to sudden and relatively large energy releases and displace-
ment increments that are not properly reflected in numerical mod-
els.
The impact of ground shaking can be assessed, but as for static
conditions, the failure related deformations are underestimated and
effects of ejection due to energy transfer cannot be easily simulat-
ed in numerical models. As a consequence, rockburst support
design methods have to largely rely on a combination of stress
analyses, empirical assessments of the depth of failure, the antici-
pated rock mass bulking, and released energy release.
Draft manuscript – Copyright protected – Cai and Kaiser 2018
112 Design principles and methodology

3.4.4 Observational design


Although rational design methods are proposed in this book and
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

numerical methods are advocated to assist in some aspects of the


design process, a rock support system design for underground
excavations should not be exclusively based on such engineering
procedures. Due to uncertainties in design parameters and support
system characteristics, any support solution will have to be scruti-
nized by ongoing verification against field monitoring and practi-
cal experience.
The authors do not agree with Stacey (2016) who suggests that
“conventional design of rock support for the containment of rock-
burst damage is not possible since neither the demand that is
generated, nor the capacity of support systems, are known, …”.
Whereas we agree that a risk-consequence approach is helpful to
confirm a design, we strongly advocate that all engineering deci-
sions must be based on a rigorous engineering design approach
that is then verified and confirmed or modified based on field
observations.
This observational design approach (Peck 1969) utilizes monitor-
ing as an integral part in the design process. The underlying logic
is that a design is not complete until the design assumptions have
been validated and the structure's performance has been matched
with performance predictions (Kaiser 1995). However, because
‘design events’ are used for support selection in burst-prone
ground, it must be realized that observations should always fall on
the safe side of predictions. This does not imply that there is an
undue safety margin.
Because all existing design methods (analytical, empirical, and
numerical) provide crude approximations of reality, it may make
sense to compare results from several design methods. However,
the authors do not advocate the indiscriminate simultaneous use of
various, particularly empirical, methods as they may not be appli-
cable or ignore the most relevant design parameters (e.g., Q con-
siders stress whereas RMR does not; numerical continuum models
do not consider bulking whereas DEM models may exaggerate,
etc.).
The bottom line is that all support designs need to be verified and
updated based on field observations and related back-analyses.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 113

3.5 Overview of rockburst support design process


The goal of a rockburst support is to mitigate the potential conse-
quences of a rockburst. For this purpose, a rock support is de-
signed to meet load, displacement, and energy demands with
appropriate support capacities under given ground and excavation
conditions. These design components are discussed briefly in this
section to introduce important concepts for the detailed engineer-
ing analyses presented in the following chapters.

3.5.1 Data collection


The first step in choosing a rock support design is to collect engi-
neering, geological, and geotechnical data based on surface, bore-
hole, and underground exploration. Design input parameters are
obtained through field mapping, logging, monitoring, and rock
mass characterization with their variability quantified. In-situ and

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
mining-induced stresses are obtained either from a stress meas-
urement program or from back-analyses of an excavation damage
or a deformation response.
A lack of geotechnical information at the early stages of a mine
design and development is usually an unpleasant fact. Hence, rock
support design at the feasibility study stage is often based on
rough estimates of the structural geology and rock mass properties.
Because rockburst problems often appear at a later stage of mining
when the extraction ratio is high, geotechnical data need to be
supplemented during the early stages of mining and should be
continuously updated. Furthermore, data on observed failure
processes and related seismicity have to be collected and docu-
mented in a systematic manner such that designs can be verified
on past experiences with excavation and support damage.
For rock support design in burst-prone ground, data collection
must include a systematic collection and analysis of historical
seismic data as discussed in detail in Volume III. The data are
needed to arrive at representative ground motions (scaling parame-
ters) from possible or likely seismic event magnitudes and loca-
tions.
Because a stress analysis may be required in a design, data on
mine layout, opening geometries, and mining sequence need to be
collected.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


114 Design principles and methodology

3.5.2 Zoning – design domains


Conditions in a mine may vary between mining blocks with dif-
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

ferent geological settings or different mining geometries. For the


design of rockburst support, the mine may have to be subdivided
into discrete support design domains, within which the key engi-
neering parameters are reasonably constant. Furthermore, zoning
is required to identify seismic design domains based on measured
or anticipated seismic activities, which are influenced by mining
activities. Within each domain, sub-zones are identified with
comparable key engineering parameters for the support design
approach illustrated in this book.
Zoning by excavation vulnerability
Because not all excavations are equally vulnerable to various
rockburst damage processes (strainbursting, shakedown, etc.), it is
most appropriate to use indicators of excavation vulnerability to
establish domains of equal rockburst potential and rockburst dam-
age severity. Means to assess excavation vulnerability in a sys-
tematic manner is presented in Chapter 1 of Volume II. Contrary
to the zoning described above based on geotechnical and seismici-
ty parameters, zoning by excavation vulnerability will assist in
fine-tuning locations of varying support requirements. From an
operational perspective, excavation vulnerability zoning may be
the most relevant approach as it is rarely possible to apply a burst
resistant support on a mine-wide or mining-block-wide basis.

3.5.3 Evaluation of support demands


One important step in rock support design is to identify potential
failure modes as explained in detail in Volume II (Figure 1-1).
Only when the anticipated failure modes are correctly identified,
can the most appropriate support design methodology be applied.
In each design domain, the load, displacement and energy de-
mands on the rock support are calculated individually considering
all possible rockburst damage mechanisms. This is achieved by
evaluating the rockburst damage severity, in terms of depth of
failure, rock mass bulking, the dynamic stress increments and
anticipated rock impact velocity (if applicable). It is often difficult
to know in advance which type of rockburst damage mechanism
will eventually dominate. Hence, it is advisable to analyse all
reasonably possible damage mechanisms and then identify the
critical support demands, i.e., worst-case but possible scenarios.
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 115

Once the demands on the support are identified, it can be evaluat-


ed whether a rock support system can be designed to prevent or
control the failure process and contain the damage. If excessive
demands are identified, other means of rockburst management
such as destressing may have to be considered to help reduce the
support demand.

3.5.4 Evaluation of rock support system capacity


Next, available rock support components are examined to identify
the best combination of support elements for an integrated rock
support system with individual support capacities exceeding the
calculated load, displacement, and energy demands. The support
selection for rockburst conditions is based on the load–
displacement characteristics of the individual support components
as well as the entire support system, consisting of compatible

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
support elements to provide rock retention, reinforcement, holding,
and connection functions. The impact of the installation sequence
on the differential loading or straining of support components has
to be considered and all possible weak links in the rock support
system are to be eliminated.
Unfortunately, there are no established methods to assess the capacity of
an integrated support system capacity. Hence, a means to estimate the
load, displacement, and energy capacity of a support system is proposed
in Chapter 4.8. For this purpose, it is assumed that all bolts work in
parallel, meaning that they are simultaneously loaded (direct loading) or
strained (indirect loading). Because bolts are not all installed at the same
time, the bolts are differentially deformed and the installation sequence
has to be considered. For the sake of simplicity, the load–displacement
characteristics of individual support components are approximated by an
equivalent perfectly plastic model to generate the cumulative support
system load, the displacement profile as well as the energy dissipation
profile of the support system.

3.5.5 Support capacity consumption and restoration


The effectiveness of support systems can be compromised by
quality deterioration (e.g., corrosion; not covered in this book) and
by the consumption of a support system’s displacement and ener-
gy capacities. In this volume, the focus is on the support compo-
nent and support system capacities and guidance is provided on
how to estimate the remnant support system capacity after some
capacity has been consumed by rock mass deformation.
Draft manuscript – Copyright protected – Cai and Kaiser 2018
116 Design principles and methodology

Support system capacity consumption


Static and dynamic stress changes deform the supported rock mass
and each support component after installation and this gradually
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

consumes support capacity. This is schematically illustrated by


Figure 3-8. If a base design is installed and then deformed to a
wall deformation d1, the support system has reached its yield
capacity and its elastic load capacity has been consumed (the
elastic energy E1 = Ee is used up). The remnant displacement
capacity to the first point of the support system degradation at d3 is
(d3 – d1 ) or the remnant energy capacity to the capacity degrada-
tion is (E2 + E3 ). Then, if during mining the support is further
deformed to d2, the remnant displacement capacity to the capacity
degradation drops to (d3 – d1), and the corresponding remnant
energy capacity is reduced to E3. In this manner, the support ca-
pacity is consumed as it is deforming.

1 2 3 4

E1 E2 E3 E4
d1 d2 d3 d4

Figure 3-8 Schematic support system characteristics illustrating four stages of support
capacity consumption [1] to [4]. Energy E1 is the energy used to deform the support from
0 to d1, E2 from d1 to d2, etc.
After the displacement d3, the support system will start losing its
load capacity. Two degradation scenarios are shown in Figure 3-8
by the blue and red support degradation curves. Even though the
retention system may at least locally fail, the overall system still
has some load and displacement capacities (remnant energy capac-
ity is the area under the red or blue curve). For example, at d4 the
remnant load capacity is 30 and 60% for the two scenarios respec-
tively, but the corresponding remnant energy capacities are highly

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 117

variable between approximately 15 (red) and 85% (blue) of E3.


For this reason and for design purposes, the displacement capacity
d3 is defined as the ‘allowable’ displacement capacity of the sup-
port system (d3 = dult).
This example illustrates that the support system capacity is pri-
marily consumed by mining-induced displacements. These dis-
placements can be imposed in a static or gradual manner or, in
seismically active mines, by co-seismic deformations. This is
discussed in more detail in Section 4.8.
Capacity restoration – PSM (Proactive Support Maintenance)
Once some support capacity has been consumed, support capacity
can be restored by proactive support maintenance (PSM), i.e., by
adding bolts offering an extra displacement capacity. In his man-
ner, the ultimate displacement capacity of the support system may,

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
for example, be increased from dult = d3 to d4 or more. If a PSM
were conducted at d2 to increase dult to d4, the remnant energy
capacity would be increased from E3 to (E3 + E4) (under full load
capacity / black line). A detailed discussion about the PSM ap-
proach on how to deal with rock support capacity selection for
ground experiencing a heavy static and repeated dynamic defor-
mation can be found in Section 4.8.

3.5.6 Design verification and modification


Support design is an ongoing process. When a rock support sys-
tem design has been chosen, it needs to be continuously assessed
for its suitability by field observation and monitoring (see Sections
3.3.7 and 3.4.4). Analyses of seismic data and failure processes
may indicate that the ‘design event’, including seismic event
magnitude or location, may need adjustment. Analyses of conver-
gence data and depth of failure observations may suggest that the
adopted rock mass properties or in-situ / mining-induced stresses
need modification. Finally, an observation of a rock support sys-
tem performance may show that the selected support components
are not compatible or performed poorly. These issues need to be
addressed promptly. More detailed discussions on these topics are
offered in Volume III.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


118 Design principles and methodology

3.6 Mitigation of rockburst damage caused by


excavation failure and dynamic disturbances
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

In this book, a distinction is made between damage caused by


excavation failure and damage caused by dynamic disturbances
from remote seismic events. For the former, rockburst damage is
not or only vaguely related to the intensity of remote seismic
events. It is dominated by the excavation vulnerability and the
availability of strain energy stored in the rock mass surrounding
the strainburst volume. For the latter, rockburst damage is linked
to the seismic stress waves and the ground motions as well as the
excavation vulnerability. In this case, it is necessary to assess the
seismic hazard to determine the PGV for design purposes.
The design methodology presented here consists of five design
components (Figure 3-9):
1) Rockburst hazard assessment (self-initiated or seismically trig-
gered strainburst or dynamically loaded rockburst);
2) Estimation of demand on support;
3) Determination of support component and support system capaci-
ties;
4) Selection of appropriate support system by fitting support ca-
pacity to anticipated demand;
5) Design verification and modification through field monitoring.
Although the overall design approach consists of only five main
components, the detailed steps involved in the design process are
complex, requiring various iterations and comparisons of alterna-
tives. For example, the failure mode and rock mass response may
be affected by the installed support requiring a revision of design
parameters depending on the selected support system. Detailed
guidance for design analyses is presented in Volumes II and III.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 119

(1) Rockburst hazard

Damage caused by excavation failure: Damage caused by dynamic disturbance


Self-initiated or seismically triggered from remote seismic event
strainbursts
Seismic hazard and ground motion

(2) Demand on support


Rockburst damage mechanisms, potential and severity
Bulking due to fracturing with or without rock ejection, and
seismic shakedown

(3) Support component capacity


Load, displacement, and energy

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
(4) Support system capacity and
support selection
Support system
capacity adjustment

(5) Field verification

Figure 3-9 Work flow for support selection in burst-prone ground.

3.7 References
Barton, N.R., Lien, R., and Lunde, J. 1977. Estimation of support
requirements for underground excavations. In 16th US Symp. Rock
Mech. Edited by F. C. and S.L. Crouch. pp. 164-177.
Bieniawski, Z.T. 1984. Rock mechanics design in mining and tunneling.
A.A. Balkema. p. 272.
Cai, M. 2013. Principles of rock support in burst-prone grounds.
Tunnelling and Underground Space Technology 36(6): 46-56.
Cai, M., and Champaigne, D. 2009. The art of rock support in burst-
prone ground. In RaSiM 7: Controlling Seismic Hazard and
Sustainable Development of Deep Mines. Edited by C.A. Tang.
Rinton Press. pp. 33-46.
Cording, E.J., Hendron Jr., A.J., and Deere, D.U. 1971. Rock
engineering for underground caverns. In Proc. ASCE Symp. on
Underground Rock Chambers. pp. 567-600.
Hoek, E., and Brown, E.T. 1980. Underground excavations in rock.
Institution of Mining and Metallurgy, London. p. 527.
Hoek, E., and Brown, E.T. 1997. Practical estimates of rock mass
strength. Int. J. Rock Mech. Min. Sci. 34(8): 1165-1186.
Hoek, E., Kaiser, P.K., and Bawden, W.F. 1995. Support of
Underground Excavations in Hard Rock. A.A. Balkema. p. 215.
Draft manuscript – Copyright protected – Cai and Kaiser 2018
120 Design principles and methodology

Kaiser, P.K. 1995. Observational modeling approach for design of


underground excavations. In Proc. Int. Workshop on Observational
Method of Construction of Large Underground Caverns in Difficult
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Ground Conditions, Tokyo. pp. 1-17.


Kaiser, P.K. 2017. Excavation vulnerability and selection of effective
rock support to mitigate rockburst damage. In Rockburst:
Mechanism, Monitoring, Warning and Mitigation. Edited by X.-T.
Feng. Elsevier. pp. 473–518.
Kaiser, P.K., Diederichs, M.S., Martin, C.D., Sharp, J., and Steiner, W.
2000. Underground Works in Hard Rock Tunnelling and Mining. In
Keynote lecture at GeoEng2000. Technomic Publishing Co.,
Melbourne, Australia. 1 pp. 841-926.
Kaiser, P.K., Tannant, D.D., and McCreath, D.R. 1996. Canadian
Rockburst Support Handbook. Geomechanics Research Centre,
Laurentian University, Sudbury, Ontario. p. 314.
Mathews, K.E., Hoek, E., Wyllie, D.C., and Stewart, S.B.V. 1981.
Prediction of stable excavations for mining at depth below 1000
metres in hard rock. CANMET Report DSS Serial No. OSQ80-
00081, Ottawa, Dept. Energy, Mines and Resources. p. 39.
Palmstrøm, A., and Broch, E. 2006. Use and misuse of rock mass
classification systems with particular reference to the Q-system.
Tunnelling and Underground Space Technology 21(6): 575-593.
Peck, R.B. 1969. Advantages and limitations of the observational method
in applied soil mechanics. Geotechnique 19(2): 171-187.
Potvin, Y. 1988. Empirical open stope design in Canada. University of
British Columbia. Ph.D. thesis, Dept. Mining and Mineral Processing.
p. 350.
Rziha, F. 1874. Lehrbuch der gesamten Turmelbaukunst, Berlin. p.
Shi, G.-H. 1992. Discontinuous deformation analysis: a new numerical
model for the statics and dynamics of deformable block structures.
Engineering computations 9(2): 157-168.
Stacey, T.R. 2016. Addressing the Consequences of Dynamic Rock
Failure in Underground Excavations. Rock Mechanics and Rock
Engineering 49(10): 4091-4101.
Sun, T. 1963. The Art of War. Translated by Samuel B. Griffith. New
York: Oxford University 65.
Terzaghi, K. 1946. Rock defects and loads on tunnel supports. In Rock
Tunneling with Steel Supports. Edited by R.V. Proctor and T.L.
White. 1 pp. 17-99.
US Army Corps of Engineers. 1980. Rock Reinforcement, Engineering
and Design, Engineer Manual, No. 1110-1-2907. p. 202.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 121

Chapter Four
Capacity of support
components
Synopsis
This chapter presents a substantially updated version of Chapter 4 in
Kaiser et al. (1996). The concept of four support functions is largely
retained and the authors made an effort to include more recent
developments in support technology and recent test data. The reader is
referred to the 1996 Canadian Rockburst Handbook (CRBSHB) for
sections that are still valid but not included here to minimize the chance
of duplication. The adopted terminology and nomenclature are

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
summarized in Appendices A and B of this volume.
First, load–displacement relations of rock support elements and factors
affecting the support capacity are presented. This is followed by a review
of laboratory and field-testing methods used to determine static and
dynamic properties of rock support components. A comprehensive
compilation of load, displacement, and energy capacities of rockbolts
and surface retaining elements is then presented. The collected data, with
all their intrinsic imperfections, form the basis for the support capacity
side of the design process. Finally, a summary of recommended
properties for engineering design is offered and detailed technical
information sheets for rock support elements are included in Appendix E
of this volume.

4 Capacity of rock support components ........................................ 123


4.1 Characteristics of rock support elements............................. 123
4.1.1 Load–displacement relation and support capacities ..... 124
4.1.2 Factors influencing support capacity ........................... 125
4.2 Rock support element testing – pull and drop tests.............. 129
4.2.1 Static pull tests ........................................................... 130
4.2.2 Shear loading of bolts ................................................. 132
4.2.3 Mesh loading.............................................................. 134
4.2.4 Laboratory dynamic drop tests.................................... 137

Draft manuscript – Copyright protected – Cai and Kaiser 2018


122 Rock support capacity

4.3 Rockbolt test results ........................................................... 143


4.3.1 Static tests................................................................... 143
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

4.3.2 Dynamic tests ............................................................. 143


4.4 Test results of surface support components ......................... 146
4.4.1 Static tests................................................................... 146
4.4.2 Dynamic tests ............................................................. 148
4.5 Summary of rock support component capacities.................. 152
4.5.1 Reinforcement and holding components ...................... 152
4.5.2 Surface retaining components ..................................... 154
4.6 Suggested design capacities for support design ................... 157
4.6.1 Design capacities of reinforcing and holding components
in burst-prone ground ................................................................ 157
4.6.2 Design capacities of retention components .................. 161
4.6.3 Design capacity compatibility ..................................... 163
4.7 References .......................................................................... 167

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 123

4 Capacity of rock support components


A rock support system is formed by the integration of various support
components such as rockbolts, mesh, shotcrete, straps, and steel arches.
Together, they interact with the host rock, fulfilling the support functions
of reinforcing the rock mass, thereby strengthening it, retaining stress-
fractured rock, preventing unraveling, and holding packages of
reinforced ground (called gabions) in place (Figure 3-5).
Unsatisfactory performance of rock support systems can often be
attributed to a lack of understanding the roles of support and the
characteristics of individual support components. A proper rock support
design must address deficiencies stemming from shortcomings and
incompatibilities of components in the system. The characteristics of
support components are addressed in this chapter and those of support
systems are featured in Chapter 5.

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
4.1 Characteristics of rock support elements
Kaiser et al. (1996) identified three characteristics of individual support
elements based on the load–displacement behaviour of support elements.
These key characteristics, stiff versus soft, strong versus weak, and
brittle versus ductile (or yielding) are illustrated by Figure 4-1a to c.
They are also applicable for integrated support systems. Each pair of
opposite characteristics highlights possible extreme choices that a
designer can make.

Strong
Load
Load

Stiff
Weak

Soft
(a) (b)
Deformation Deformation
Brittle Lp
Lult
Load
Load

Ductile/yielding
Pre-load
Ep Ey

(e)
dp dult
(c) (d)
Deformation Displacement

Figure 4-1 Key characteristics of support elements.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


124 Rock support capacity

Each support function (reinforce, retain, hold, and connect) can be


fulfilled by support elements with these characteristics. For example,
retaining elements can be either stiff and strong (e.g., closed ring of
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

shotcrete) or soft and ductile (e.g., chain-link mesh or slotted shotcrete


panels). For rockburst support, the desired characteristics of support
elements are stiff and strong and ductile or yielding. These
characteristics are highlighted in red and bold in Figure 4-1. Stiff means
that a support element is attracting load at small displacements. Strong
support elements have high load capacities that can hold large masses of
massive or broken rock. Yielding support elements have high
displacement capacities to accommodate large rock deformations. Some
yielding support elements have a steady load capacity over a large
displacement range such that they can absorb much energy. A rockbolt
that possesses these three qualities (stiff, strong, and ductile) is a superior
bolt for burst-prone ground. Similarly, a rock support system that
seamlessly integrates all three qualities is a superior support system.
Unfortunately, it is difficult to assemble these desired characteristics in a
single support element type. For example, yielding support elements are
often not stiff. Hence, one needs to combine different support element
types to form a system that exhibits the desired overall characteristics
(Chapter 5). The most critical aspect of support selection is to ensure full
compatibility of all support elements in a support system. It works best if
all the support elements in the system work in harmony.

4.1.1 Load–displacement relation and support capacities


Figure 4-1d shows a generic load–displacement curve for a
reinforcing/holding element. The maximum or peak load Lp and the
ultimate load capacity Lult define the load capacity. Lp and Lult need to be
maintained during yielding (for yielding bolts) or before failure when the
support element becomes ineffective. The corresponding displacements
are dp and dult at the peak and ultimate load, respectively.
The energy absorption capacity is a function of the load and
displacement capacities, the pre-loading, the loading rate, and the shape
of the load–displacement curve. The pull-out energy absorption
capacities are equivalent to the work performed during a pull or direct
impact test, and it can be calculated by integrating the area under the
load–displacement curve. The available energy absorption capacity after
preload at the peak load is Ep, and the total, ultimate energy absorption
capacity is Eult = (Ep + Ey).
Typical static direct pull-out test results are presented in Figure 4-2
(Kaiser et al. 1996) for reinforcing and holding elements. Dynamic test
results are presented in Section 4.3.2 and properties for other currently
available support components are discussed and summarized in Table C-
1 in Appendix C.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 125

(a)

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
(b)
Figure 4-2 Direct loading tests: (a) load–displacement curves and (b) static
energy absorption capacities for reinforcing and holding elements (Kaiser et al.
1996).

4.1.2 Factors influencing support capacity


Loading mode
The capacity of individual support components depends on the mode of
loading: ‘direct’ at the plate or ‘indirect’ by opening of fractures inside
the rock mass. Published support characteristics (e.g., as shown in Figure
4-2) and quality assurance test data are typically obtained from direct
pull-out tests.
As discussed in Chapter 3, bolts are not necessarily or exclusively loaded
directly at the plate; they may also be strained by a push-action from
inside the failing rock mass. The load–displacement curves for indirect

Draft manuscript – Copyright protected – Cai and Kaiser 2018


126 Rock support capacity

loading obtained from split pull tests differ from direct test results
(Figure 4-2). Whereas the load capacity is generally maintained, the
displacement capacity is increased by the dual pull-out action. For some
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

bolt types (e.g., rebar) it may be as much as double. The energy


absorption capacity may therefore, at locations of indirect loading,
locally be much (up to two times) higher due to the two-sided pull-out
displacement. The characteristics presented in Figure 4-2 provide lower
bound values for dult and Eult.
If a bolt is allowed to simultaneously yield at the plate (direct loading)
and internal to the rock mass (indirect loading), its energy dissipation
capacity is about (1 + 2n) Eult, where n represents the number of internal
bolt yield locations. If yielding at the plate is prevented, the internal
energy dissipation capacity is about 2nEult. Because it is impossible to
predict a priori at how many locations a bolt will actually dissipate
energy, it is reasonable to assume that the energy dissipation capacity is
between one and two times the Eult value as shown in Figure 4-2b. The
capacity of rock reinforcements is likely much higher than that inferred
from direct pull-out tests.
Loading rate
The load and energy capacities of a support element obtained from the
static tests may not be representative of the actual dynamic load and
energy absorption capacities because the shape of the load–displacement
curve depends on the loading rate. Static tests may both over- and under-
estimate the dynamic capacity. Whereas steel generally has a higher
capacity at high loading rates, the ultimate displacement capacity may be
reduced with the consequence that the energy dissipation capacity is
lowered.
Load sharing between support components
In practice, pre-stressing and the installation sequence influence the
available displacement and energy capacities. Some of the deformation
capacity of individual components may be used up due to pre-stressing
or mining-induced straining because not all components are deformed
simultaneously. Hence, the actual load, displacement and energy
capacities of each component depend on the remnant displacement
capacity after pre-loads or pre-deformation have been applied. The
remnant load capacity is zero once the yield point has been reached, i.e.,
no additional weight can be added when a bolt is yielding. The remnant
displacement capacity is the difference between the ultimate and the
previously imposed displacements. Accordingly, the remnant energy
capacity is the variance between the total and the previously consumed
energy capacity.
This is illustrated for an example with stress-fracturing in the roof of a
tunnel supported by a system of point-anchored rebar, Split sets (or
friction sets) and conebolts connected to mesh/straps. This combination
of bolts was deployed in an sequential manner at Beaconsfield Mine in
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 127

Australia (Scott et al. 2008). This eventually led to mining-induced roof


collapses because the combined load capacity gradually decreased with
increasing deformations during stress-fracturing. In practice, the three
bolt types were not installed simultaneously but, for illustrative purposes,
it is assumed that they were all installed at the same time.
As mining progressed, the stress in the roof of the drifts first rose and
then dropped as the depth of stress fracturing increased rapidly (Figure
4-3a). Because of the associated rock bulking (Kaiser et al. 1996), the
deformation of the back and thus the bolts increased rapidly (Figure 4-3b)
in the post-peak deformation stage.

Rock mass bulking


Pillar edge stress

2-10%

1-2%
Rock mass strain

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Deformation of
Depth of failure

drift back

1-2 m 2-20 cm

(a) (b)

300

250

200

Resin-grounted rebar
Load (kN)

Conebolt
150

100 Swellex bolt


Mechanical rockbolt
Split set

50

0
0 50 100 150
Displacement (mm)

(c)
Figure 4-3 Schematic illustration of the remnant holding capacity of a support
system consisting of point-anchored rebar, Split sets and conebolts connected to
mesh/straps: (a) and (b) strain-controlled failure in back; (c) related holding
capacity of support system.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


128 Rock support capacity

The cumulative load capacity is illustrated by Figure 4-3c, assuming that


the load–displacement curves of Figure 4-2a for direct loading are
applicable and that all bolts were installed at the same time. At 30 mm,
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

all three types of bolts are near their peak load bearing capacity for a
total capacity of about 70 + 110 + 120 = 300 kN; however, at 50 mm the
point-anchored bolt has failed at the plate and the combined capacity
drops to about 110 + 120 = 230 kN. At 100 mm, the depth of failure is
such that the anchor length of the Split set becomes insufficient to hold
more than a few tons, the combined capacity drops to that of the
conebolts alone or about 150 kN. Most importantly, when one bolt type
is eliminated (fails), the overall bolt spacing is drastically increased,
eventually to that of conebolts alone. As a consequence, the capacity of
the retention system is reduced (or reduces) simultaneously.
Allowable support system displacement
The allowable displacement of a support system can be defined in two
ways: (1) by what is allowable from an operational perspective (e.g.,
acceptable drift closure), and more importantly, (2) by the displacement a
support system, composed of various support components, can sustain
without losing the integrity of the combined support system.
The allowable displacement capacity of a support system is typically
much less than the displacement capacity of the most deformable support
component but it may also be higher than the ultimate direct loading
displacement capacity of the least deformable support component. For
example, a rebar may have as little as 15 to 20 mm of ultimate pull-out
displacement capacity; however, when integrated into a support system it
will be able to deform the same amount near the plate as well as 30 to 40
mm (indirect loading) displacement at several distances from the wall
(inside the rock mass). As long as other support elements ensure the
integrity of the support system and distribute the rock mass strain
without causing strain localization, an apparently brittle bolt (rebar) will
survive and its allowable deformation is much larger than that implied
from pull tests.
Cumulative energy capacity of a support system
The cumulative energy capacity of a support system depends on the
installation sequence and the mining-induced deformation that each
support component has experienced. For example, for a support system
where all components were simultaneously installed before mining and
by reference to Figure 4-2b, at a displacement of 100 mm (if applied at
the wall), a rebar and a mechanical bolt would have failed (but each
would have dissipated about 4 kJ before failure) and all other bolt types
would have dissipated between 7 and 13 kJ. Therefore, if the support
integrity at this displacement threshold is maintained, the combined
energy dissipation capacity of a yielding SwellexTM bolt installed in
parallel with a rebar would be 10 + 4 = 14 kJ, not just its capacity of 10
kJ. This example illustrates that the ultimate energy capacities at dult,
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 129

ranging from 10 to > 30 kJ for individual components (Figure 4-2), may


not be of practical relevance because not all bolts can simultaneously
reach their ultimate displacement capacity dult.
This consideration introduces a relevant design constraint with respect to
practically available displacement and energy dissipation capacities. The
simultaneous capacity of a support system with various support
components gradually increases with increasing deformation but rarely
reaches the cumulative ultimate energy dissipation capacity. For this
reason, we will later introduce the concept of E100 or E200, i.e., the energy
dissipation capacity of a support component at an allowable
displacement threshold of 100 or 200 mm, respectively.
Remnant energy capacity
If a support component or support system has been pre-loaded or pre-
deformed by mining-induced straining, the remnant displacement and
energy capacities of the individual support component and thus the
capacity of the support system is reduced. For example, if a support is

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
designed for an allowable displacement threshold of 200 mm but it has
already experienced 100 mm of displacement, the remnant displacement
capacity is only 50% of the design capacity. For a perfectly plastic
support, the remnant energy dissipation capacity would therefore also be
reduced to 50%.
In burst-prone mines, it is not the installed capacity but the remnant
capacity available at the time of bursting that matters. If too much
capacity has been consumed at a given mining stage, it may be necessary
to proactively restore the support system’s displacement capacity and
thus its energy dissipation capacity.

4.2 Rock support element testing – pull and drop


tests
Various test methods have been developed to determine the capacities of
support element and support systems. The test methods can be classified
into static and dynamic testing methods based on the type of loading,
laboratory and in-situ testing methods based on test location, and single
element and support system testing based on test arrangement (Figure
4-4). In a dynamic test, the dynamic load is applied in the form of a drop
weight impact, moment transfer, stress waves generated from blasts, or
stress waves generated from actual rockbursts.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


130 Rock support capacity

Static Pull test

Load type Drop test


Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Dynamic Simulated rockburst

Rockbolt Ground truthing


Test
Surface support
component
Support system

Pull test
Lab test
Drop test
Test
location
Pull test
Field test
Simulated rockburst

Figure 4-4 Classification of test methods to determine rock support capacity.


Test results on support components are reviewed in this chapter and
those from testing of support systems in Chapter 5.

4.2.1 Static pull tests

Indirect and direct bolt loading


The double-embedment split laboratory test method (Stillborg 1994)
provides indirect loading load–deformation characteristics of rockbolts.
Rockbolts are installed across a simulated ‘joint’, using two blocks of
high strength reinforced concrete. The load is applied to the bolt by
separating the two concrete blocks (Figure 4-5a). In dynamic test
systems, the equivalent approach uses split tubes.
The most frequently applied bolt test method is conducted by force
application at the plate or the thread (direct loading tests). This produces
results as presented in Figure 4-2. A direct pull-out test can be conducted
in the laboratory by installing the rockbolt into a steel pipe (Figure 4-5b)
or a concrete block. It can be conducted in the field as shown in Figure
4-5c.
For a grouted rockbolt loaded at the plate, the free (ungrouted) stretch
length is Lfs, and the applied bolt force F drops to zero at a distance of Ly
= Lg + Lfs, where Lg is the gradual deformation zone length in the grouted
section. For an indirectly loaded grouted rockbolt (e.g., split-tube at a
joint), the applied bolt force F drops to zero at a distance of Ly = Lg + Lfs
on both sides. Thus, the total yield length is Ly = 2(Lg + Lfs).
Most rockbolts have a threaded section near the collar for tightening the
plates. Some bolts such as forged-head rebar do not have threaded
sections and these bolts are used in certain applications where tightening
the plates is not required. The tensile load capacity F is often smaller at
the thread than in the main shaft of the bolt. When a bolt at the threaded
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 131

section breaks prematurely, its elongation potential cannot be fully


realized by direct loading. For this reason, the double-embedment split
test method (or indirect split-tube test) tends to give a slightly higher load
capacity and a much larger straining capacity for the same bolt type
when compared with direct pull-out test results. If the free stretching
length Lfs is zero, as would typically be the case at an open joint or
fracture, the peak and the ultimate load capacities of both loading types
are comparable. The ultimate displacement capacity almost doubles for
the indirect loading test.

Rockbolt
Concrete block

F
Lg 2Lfs Lg

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
(a)
F

Resin grout
Lg Lfs
Steel pipe
(b) Rockbolt

(c)
Figure 4-5 (a) Laboratory double-embedment split test method; (b) laboratory
direct pull test method; (c) field direct pull test setup.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


132 Rock support capacity

Strainburst loading of bolts


Two loading modes are encountered during strainbursts: internal
straining of the bolt shaft by fractured and bulking ground, and loading
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

of the bolt head by transfer of load from the retention system to the plate.
The split-tube (indirect) test method provides the load–displacement
characteristics of a bolt crossing a joint or a fracture. This type of loading
is likely to occur inside the strainburst volume at some distance from the
wall. Hence, double-embedment split tests are more representative for
understanding the capacity of rock reinforcements and bolts that are
yielding at some distance from the excavation wall.
The (direct) pull test method provides the load–displacement
characteristics of a bolt when loaded via the plate. This type of loading
occurs at the wall where broken rock impacts the retention system. An
example for a load transfer from retaining elements to the plate and then
to the bolt is illustrated in Figure 2-16 . Evaluating the load capacity by
direct pull test at the collar is therefore representative for situations
where bulking occurs between the bolts.
For these reasons, extreme care must be exercised when selecting design
parameters or using test results for support design.

4.2.2 Shear loading of bolts


There is one major deficiency in these pull tests, i.e., the bolts are tested
in perfect, axial tension. In practice, bolts can also be sheared,
particularly when shear rupture localization occurs. Whereas various bolt
types have been tested in shear or in combined tension and shear (e.g.,
Bjurstrom 1974; Aziz et al. 2003; Snell et al. 2017), most bolt capacities
published in the rockburst literature do not consider the capacity
reducing effects of shear.
As discussed in Chapter 2, shear rupture is often associated with
strainbursts and shear localization is to be expected when large wall
displacements are encountered due to rock mass bulking. Unfortunately,
there are no systematic dynamic shear test data available. However, from
static tests on anchors and dowels, it is known that, depending on the
steel’s ductility, the pure shear capacity is between 58 and 75% of the
pure tension capacity at the yield point and between 75 and 90% of the
steel’s ultimate strength. The shear resistance decreases with increasing
tensile load as illustrated by Figure 4-6.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 133

120

16 mm
100 22 mm
double 16 mm bolts

Shear strength (kN)


80

60

40

20

0
0 50 100 150 200
Tensile yield strength (kN)

Figure 4-6 Static shear versus axial load capacity diagram at yield stage assuming
58% shear strength at zero tension.

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Figure 4-6 illustrates an often-ignored interdependence of shear and
tensile strengths:
(1) in pure shear without tension, the shear capacity is between 58%
of the tensile capacity for standard steel at yield and 90% at
ultimate for ductile steel;
(2) this shear capacity is nonlinearly reduced with increasing tensile
stress and reaches approximately half of the above quoted shear
capacities when the axial stress reaches about 85% of the steel’s
tensile strength; and
(3) when steel is yielding in tension, the shear resistance is zero.
The first situation (1) is rarely encountered in the field because there is
almost always some axial loading in bolts. Hence, it is reasonable to
assume that the shear resistance is at best 50% of the pure tension
capacity.
The second situation (2) represents the status of a bolt that has not
reached the yield point, e.g., a conebolt with the cone sliding through
grout. Because it is reasonable to assume that at least 85% of the axial
capacity is reached when a cone is ploughing in the grout, the remnant
shear capacity of a conebolt is less than 25% of the capacity in pure
tension.
The third situation (3) is encountered in bolts that depend on steel yield
(e.g., the D-bolt). In this case, there is essentially no shear capacity
available at locations where the bolt yields. The same applies locally for
conebolts when the grout is too strong or debonding is not effective and
some plastic steel shaft deformation is experienced.
The practical implication is that:
Draft manuscript – Copyright protected – Cai and Kaiser 2018
134 Rock support capacity

- it is not conservative to design support on axial capacities alone;


reductions for shear need to be considered; and
- steel with axial loads of less than about 85% of yield strength is
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

required to provide a reliable dowelling effect.


It is stated earlier that stiff rebars are needed to control and minimize
rock mass bulking. The need for shear resistance when shear localization
or shear rupture is anticipated provides a further justification for the use
of ‘un-tensioned’ rebar in burst-prone ground. Even if a rebar is yielding
(and possibly failing near the plate), it remains unstressed beyond the
stretch length (of typically < 0.5 m) unless severe rock mass bulking
loads the bolt inside the rock mass. For strainburst-prone ground, it is
paramount that ‘un-tensioned’ rockbolts are provided in the burst volume.

4.2.3 Mesh loading


The load–deformation characteristics of surface support elements such as
mesh and shotcrete are obtained using pull tests, either in the laboratory
or in the field. A laboratory setup at GRC (Geomechanics Research
Centre, Laurentian University, Canada) is shown in Figure 4-7.
Sometimes, it is convenient to apply the load by pushing instead of
pulling (Kirsten and Labrum 1990; Morton et al. 2009).
In the test setup shown in Figure 4-7, the mesh is bolted to the frame and
the relative displacement of the bolt with the frame is almost zero. In the
field, the bolts can deform under loading, and the total displacement at
the pull plate is equal to the sum of the mesh displacement relative to the
bolts and the displacement of the bolts.

Figure 4-7 GRC laboratory test facility for static mesh and shotcrete panel testing.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 135

The GRC test facility currently allows for testing of two bolt patterns
with 0.74 and 1.49 m2 loading areas, respectively, as illustrated by Figure
4-8.

Figure 4-8 Diamond (left) and square (right) test patterns in GRC laboratory with
0.74 and 1.49 m2 test areas, respectively.
Previously published examples of mesh retention elements are

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
reproduced in Figure 4-9 (Kaiser et al. 1996). Much more work has
been completed on retention systems since then and this will be
discussed and summarized later in this book.

(a) (b)
Figure 4-9 (a) Load–displacement curves and (b) static energy absorption
capacities for some mesh retention elements (Kaiser et al. 1996).
It is important to understand that the central deflection, the load, and the
energy dissipation capacities of a mesh are highly dependent on the
adopted bolt pattern. Figure 4-11 presents test results on #6 gauge mesh
with diamond and square patterns and the results are compared in
normalized form in Table 4-1.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


136 Rock support capacity

35

30
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

25

20

Load (kN)
15

10

0
0 100 200 300 400
Displacement (mm)
(a)
35
30
25
Load (kN)

20
15
10
5
0
0 100 200 300 400 500 600 700
(b) Displacement (mm)
Figure 4-10 Load–displacement curves of #6 gauge mesh with (a) diamond bolt
pattern, and (b) square pattern.

Table 4-1 Comparison of normalized mesh capacities for tests presented in


Figure 4-1
Bolt L1st peak d1st peak Lmax dmax E1st peak Emax
pattern (kN/m2) (mm) (kN/m2) (mm) (kJ/m2) (kJ/m2)
Diamond 32 170 41 210 1.8 3.0
Square 14 370 20 525 1.7 3.8

Note: because of the 1.5 m ´ 1.5 m (5 ft ´ 5 ft) mesh size used in the tests, the
central deflections and thus the energy capacities are likely lower.

Several noteworthy observations can be made from Table 4-1. Compared


with the results of the square pattern setup, the load capacity of the mesh
with diamond pattern is more than double and the central deflection is
less than half. As a consequence, the energy dissipation capacities per
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 137

square metre are comparable. This example illustrates the practical


benefit of tight bolt spacing. Tight bolting decreases the mesh deflection
while increasing the load capacity and maintains a comparable energy
dissipation capacity.

4.2.4 Laboratory dynamic drop tests


The behaviour of many engineering materials is loading rate dependent
(Kaiser et al. 1996; Gomez et al. 2001; Cai et al. 2007). Hence, dynamic
tests are needed to determine the dynamic capacities of rock support
elements. Tests on steel alone indicate that the high loading rate capacity
of steel can be between 10% and 40% higher than the static capacity
(Malvar and Crawford 1998). However, strain localization often reduces
the ultimate displacement capacity when loaded at very high rates. As a
consequence, the dynamic energy dissipation capacity of steel for
rockburst loading rates may not be much higher than the static capacity.
Drop test facilities to perform repeated and repeatable dynamic loading
tests have been built in Canada, South Africa, Sweden, Australia,

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Switzerland, and China. They have been widely used to test the dynamic
properties of reinforcing, holding and retaining elements. Drop test
facilities allow conducting repeated loading tests at a relatively low cost.
Some argue that the performance of bolts under repeated loading differs
from those conducted by single impact to failure. While this is true, it is
incorrect to conclude that repeated tests are of no practical value. In fact,
most support systems in burst-prone mines are incrementally and
repeatedly loaded by static and dynamic displacements before reaching
the failure point.
Dynamic tests on rockbolts
The CANMET drop test facility (Figure 4-11a) in Canada, originally
designed in the late 1990s for Noranda Technology Centre (NTC) by
Maloney and Kaiser of GRC, was eventually upgraded by CANMET to a
maximum weight capacity of 3 tonnes, and a maximum input energy of
58.9 kJ at a drop height of 2 m (Doucet 2012). The facility applies a
dynamic load directly to the test object. Recently, ASTM (2008) has
adopted the NTC-CANMET system as a standard method for laboratory
determination of rock anchor capacity by drop tests. Similar drop test
facilities have recently been built at China University of Mining and
Technology in Beijing and will be built in Chile in collaboration of
MIRARCO of Laurentian University with the University of Chile.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


138 Rock support capacity
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

(a) (b)
Figure 4-11 (a) CANMET drop test facility; (b) WASM drop test facility (Player
et al., 2004)
Another type of drop test facility employs a momentum transfer
mechanism to apply a dynamic load to the test object (Ansell 1999;
Player et al. 2004; Ansell 2005; Player 2012). The WASM (Western
Australian School of Mines) facility (Figure 4-11b) has a weight capacity
of 4.5 tonnes, a drop height of 6 m, and a maximum input energy of 225
kJ. Some energy can be absorbed by the buffer system and monitoring is
used to establish the exact amount of energy transferred to the test target.
Facilities developed in South Africa, which employ different impact
loading methods, are described by Stacey and Ortlepp (1999), Stacey and
Ortlepp (2001), Ortlepp and Swart (2002), and Ortlepp and Erasmus
(2005). In the facility of Stacey and Ortlepp (1999), a swing beam was
used to receive the impact load and transfer it to rockbolts. Ortlepp and
Erasmus (2005) developed a wedge-block loading device to convert a
vertical displacement into a horizontal displacement to load Duraset
rockbolts. Among all the test rigs, the CANMET and WASM facilities
are currently the most active ones in use.
Two types of load transfer setups are used to test the dynamic capacities
of rockbolts: direct impact and double-embedment or split-tube indirect
tests. For the direct impact test, the rockbolts are installed in thick-wall
steel pipes (tubes) and the drop weight directly impacts the plate attached
at the threaded section of the bolts. In the split-tube test, a bolt is
installed in two joining steel pipes and the drop weight impacts a steel
seat welded to the lower steel pipe (Figure 4-12). A split indirect impact

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 139

hammer is under development for the University of Chile to conduct


sequential indirect impact tests along the bolt shaft.
In the direct impact test, the bolts are stretched between the nut holding
the plate and a distance into the pipe, called free stretching length Lfs, and
the gradual deformation zone length Lg that depends on the grout and
rock quality (strength) and the rod surface properties (e.g., smooth vs
rough or debonded). For D-bolts, Lfs is approximately the distance to the
first paddle anchor and for conebolts, if perfectly debonded, to the cone.
In the indirect test, double-sided pull-out is simulated and the total yield
length is 2(Lg + Lfs) for rebar, meaning that the consumable energy may
double (if all other factors remain unchanged). For fully grouted rebar,
Lfs = 0 and 2Lg of about 20 to 30 mm was measured in a set of static pull
tests using two concrete blocks (Stillborg 1994). Rebar and threadbar
may have a higher stretch length and may absorb more energy if the
grout or rock quality is lower. For conebolts the yield length is, at least
theoretically, unchanged and the performance should be comparable to
that in the direct loading test. The lower split pipe will transfer some, if

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
not most, of the load or energy to the plate attached at the threaded
section of the bolt.
For D-bolts, the yield length can be engineered by choosing the paddle
separation. For the example shown in Figure 4-12, the distance between
the two paddles (or anchors) adjacent to the split is 0.8 m. For a D-bolt,
most of the impact energy is absorbed by steel stretching between the
two paddles. The movement of the lower split pipe will transfer most of
the energy to the lower paddle but some loads or energy may also be
transferred to the plate attached at the threaded section of the bolt. If the
first anchor (near the collar) loses its anchoring capacity, most of the
load will be imposed on the plate.
It is evident that the bolt capacities determined by these two test methods
will differ and it is important to consider which dynamic loading mode is
representative of conditions encountered in the field. For example, for
failure modes by momentum transfer from rock blocks moving at a
defined velocity and if rock mass bulking between bolts leads to large
deformations of the retaining components, the direct impact test is most
representative. On the other hand, for strainbursting whereby stress-
fractured blocks are separating and moving relative to each other inside
the reinforced rock mass, the indirect test is more representative to assess
the energy dissipation capacity of the reinforced rock mass. In reality,
both loading modes co-exist.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


140 Rock support capacity
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Figure 4-12 Double-embedment split-tube drop test setup; shown for D-bolt with
three anchor/paddle points (modified from Li 2010).
Ortlepp (1994) conducted dynamic tests on rebar and conebolts using
blast loading at a quarry site in South Africa. Different rockbolt types,
grouted into holes drilled into the quarry floor, held down six identical
reinforced concrete blocks against a levelled concrete surface.
Explosives were placed between the concrete blocks and the floor to
generate dynamic uplift loading. The test results showed that both rebar
types, 16 and 25 mm in diameter, failed to resist the strain localization
due to loading at the blast location. On the other hand, the 16 and 22 mm
diameter conebolts, possessed sufficient energy absorbing capacities,
survived the impulse loading because a sufficiently long yield length was
provided. This type of dynamic test was developed to promote the
acceptance of yielding bolts in burst-prone grounds. These tests
demonstrated the effectiveness of yielding bolts to dissipate the impact
energy.
Dynamic tests on surface support
Drop test facilities have also been built to test the dynamic capacities of
retaining or surface support elements such as mesh and shotcrete panels.
The GRC facility (Figure 4-13) has a 565 kg drop-weight, with impact
velocities ranging from 4.4 to 7.7 m/s and a maximum impact energy of
16.6 kJ (Kaiser et al. 1996; Tannant et al. 1997). The South Africa drop
test facility (Figure 4-14) has a maximum drop weight of 2.7 t, a
maximum drop height of 3.3 m, and a maximum input energy of 70 kJ
(Stacey and Ortlepp 2001). More recently, a test facility (Figure 4-15)
was built in Walenstadt, Switzerland, where a 3.6 m × 3.6 m surface
support can be dynamically tested with four dynamic rockbolts in a
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 141

regular bolt pattern (Bucher et al. 2013; Brändle et al. 2017). This facility
has a drop height of approximately 3.25 m and a drop weight of 6.28 t to
provide about 200 kJ input energy. It is in many ways similar to the
South Africa drop test facility except that the rock mass (represented by a
thin concrete slab and natural rock boulders) around the bolts is bonded
to the bolts and held by the plates. As a consequence, part of the impact
load goes directly to the bolts. The facility is equipped with load cells
and accelerometers as well as two high-speed cameras such that the
energy consumption by the bolts, mesh, and test setup can be established
(Bucher et al. 2013).
The GRC facility is used to test surface support elements that are held by
stiff reinforced columns and the drop weight directly impacts the target.
This is equivalent to direct impact testing of surface support elements
installed using stiff, non-yielding rockbolts. The South Africa test facility
uses concrete blocks stacked as a pyramid within a predefined rockbolt
pattern to indirectly transfer the impact load from the drop weight to the
target. Part of the impact energy is consumed in crushing and deforming

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
the concrete blocks. A distinct feature of the South Africa and the Swiss
facilities is the tested retaining element (mesh or shotcrete panel or
composite panel) that can be held by yielding bolts, creating a softer load
transfer system compared to the GRC facility.
Recognizing the differences in loading and boundary conditions is
important when interpreting and using respective test results for a
support design. Because of the stiff boundary conditions and direct
impact loading, the energy absorption capacities obtained for surface
support elements tested by the GRC facility have lower bound values.
There is one benefit of the GRC facility; it tests the robustness of the
connection between bolt/plates and the retaining system and presents a
measure of the pure retention capacity. On the other hand, due to energy
losses (resulted from the momentum transfer via pyramid blocks or
concrete slabs) coupled with even load distributions and the influence of
the yielding holding elements, the energy absorption capacity obtained
from the South Africa and the Swiss facilities likely represent upper
bound values.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


142 Rock support capacity

deflections
measured here
supp ort

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
columns plan

shotcrete 1.2 m

565 kg guide for


0.6 m diameter drop -weight
cylindrical drop-weight
with central hole

nut and plate

shotcrete

load
cells
section
b races

steel p late 0.2 m diameter, 1.2 m high


p ip es filled with concrete

reinforced concrete p ad

Figure 4-13 GRC drop test facility (Kaiser et al. 1996); area between 4 bolts
measures 0.74 m2.

Figure 4-14 South African drop test facility (Stacey and Ortlepp 2001).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 143
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Figure 4-15 Switzerland drop test facility (Bucher et al. 2013).

4.3 Rockbolt test results


4.3.1 Static tests
The load–displacement relations obtained from pull-test on various
rockbolts are presented in Figure 4-2a and the corresponding energy–
displacement relations obtained by integrating the area under the load–
displacement curve for each rockbolt are shown in Figure 4-2b. The load,
ultimate displacement, and energy absorption capacities are discussed in
Sections 4.1.1 to 4.1.2 and are summarized in Table C-1.

4.3.2 Dynamic tests


Figure 4-16 compares the dynamic energy absorption capacities of
various rockbolt types obtained from drop tests. Data for debonded
threadbars and D-bolts were acquired by split-tube testing and all the
others were obtained from direct impact tests. The toe-anchored
threadbars (fully grouted to collar) failed at the thread and absorbed little
Draft manuscript – Copyright protected – Cai and Kaiser 2018
144 Rock support capacity

energy (< 2 kJ). Toe-anchored rebar behave similarly and their direct
loading energy absorption capacity is equally small. The debonded 20
mm diameter threadbars were debonded over a 1.6 m long section, which

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
allowed the steel to stretch in order to absorb dynamic energy. Two
lengths of free stretch sections, 0.846 m and 1.5 m between paddles,
were tested for the 22 mm D-bolts. All data points are from the first drop
weight impact. All but the toe-anchored threadbar and rebar did not fail
upon first impact.
A linear trend between steel straining (displacement) and energy
absorption capacity can be seen from data of threadbar and D-bolt tests,
because these two rockbolts rely on steel stretch (or plastic steel
deformation) to absorb dynamic energy. D-bolts with longer free stretch
sections absorb more energy. As indicated by the left swinging blue
arrow, the slope of the linear trend line showing steel straining and
energy absorption capacity will be higher if high strength steel is used to
make the rockbolts or if the bolt diameter is larger.

100% Steel stretching to absorb energy (22 mm)


60
Increasing
bolt diameter 100% Steel stretching
or steel strength to absorb energy (20 mm) 0% steel stretching
50 (100% cone plow)
Increasing
to absorb energy
steel stretching
Dynamic input energy (kJ)

40

30 MCB33 (new, 17.2 mm)

MCB33 (old, 17.2 mm)

Fully bonded threadbar (20 mm)


20
Fully debonded threadbar (20 mm)

Toe anchored threadbar (20 mm)

10 SA Conebolt (22 mm, 40 GPa grout)

D-bolt (22 mm, 0.846 m segment)

D-bolt (22 mm, 1.5 m segment)


0
0 200 400 600 800 1000

Displacement (mm)

Figure 4-16 Comparison of energy dissipation capacity of various rockbolt types.


Threadbar data are from Player et al. (2009), D-bolt data from Li and Charette
(2010) and MCB data from Cai et al. (2010); Cai and Champaigne (2012). Red
arrows are placed as reference for explanation in text.

Some steel stretch also occurs in the MCB conebolts depending on the
debonding medium and loading rate, even though the design intention is
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 145

to let the cone plough through the resin to absorb dynamic energy. Based
on the test data, a boundary between pure cone plough and a combination
of cone plough with steel stretch can be established as shown in Figure
4-16 (red solid line).
It becomes evident from the slope of the lines drawn in Figure 4-16 that
the energy dissipation rate strongly depends on the yielding mechanism
of a bolt. When steel stretching is involved to dissipate energy, high
energy dissipation rates, in terms of kJ/mm, can be achieved. For bolts
without a sliding mechanism, high energy dissipation rates can only be
realized by deploying high strength steel or by using larger diameter bars.
When a sliding mechanism is involved, the energy dissipation rate is in
general smaller but the sliding mechanism brings one major advantage as
relatively large deformations can be facilitated by the sliding of the bolt.
This is important when severe rock mass bulking occurs and large wall
displacements have to be accommodated after a rockburst, i.e., when a
large displacement capacity is desired. In other words, bolts with a
sliding mechanism do, in general, retain more energy dissipation
capacity after some displacement has been imposed. Hence, bolts with
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

sliding mechanisms tend to retain a higher safety margin in terms of


remnant energy dissipation capacity.
From an operational perspective, however, large wall displacements (>
200 to 300 mm) are undesirable (to minimize rehabilitation requirements)
and, from a support system stability perspective, large deformations put
the support system integrity at risk as retention components (e.g.,
shotcrete) may get excessively strained and shear localizations may
occur (preventing the full mobilization of the theoretical pull-out
capacities). Hence, the operational space is highlighted by the red arrows
in Figure 4-16. Upon an initial impact with 200 to 250 mm displacement1,
the bolts relying on steel stretch will have dissipated between 40 and 50
kJ whereas those with pure sliding mechanisms will only have dissipated
between 10 to 13 kJ (based on 17.2 mm diameter MCB conebolt data).
The energy dissipation rates can be higher if larger diameter bolts are
used and there is a combination of sliding and steel stretch.
Finally, it is important to differentiate the deformation source of the
rockbolts shown in Figure 4-16. Dynamic loads were directly applied to
the plates for toe-anchored threadbar and MCBs but not to the debonded
threadbar and D-bolts. Hence, the former displacement is external (direct)
and the latter is internal (indirect). The ideal yielding bolt provides a
large displacement capacity near the plate, or at the excavation surface,
and a high internal energy dissipation capacity.

1
For D-bolt, this wall displacement is not necessarily equal to the steel stretching
between two anchors.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


146 Rock support capacity

Both characteristics are required in an integrated support system, where


the yielding support elements are combined with surface retaining
elements such as mesh and shotcrete, and a reinforcement of the rock

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
mass bulking zone is needed. One without the other can lead to a weak
link scenario. Deformable, direct loading resistance must be provided at
the surface (at connections between bolts and retaining components) and
internal straining resistance must be provided to minimize bulking and to
prevent shear localization or shear rupture.

4.4 Test results of surface support components


4.4.1 Static tests
The load–displacement curves for #4-, #6-, #9-gauge weld mesh (0.1 m
wire spacing) and #9-gauge chain-link mesh are shown in Figure 4-17.
The mesh was held by four bolts in a 0.85 m spacing diamond pattern
and pull tested at the centre (for details refer to Tannant et al. (1997) and
Kaiser et al. (1996)). For the weld mesh, the peak loads were followed by
a sudden drop in load, which was caused by failure of individual mesh
wires. The full capacity of a mesh is only mobilized after substantial
deformation, roughly 110 to 150 mm for weld mesh and 420 mm for
chain-link mesh2. The chain-link mesh has more post-peak ductility than
the weld mesh. Compared with chain-link mesh, weld mesh has a high
initial loading stiffness.

Figure 4-17 Load–displacement curves of weld mesh and chain-link mesh under
static loading (modified from Tannant et al. (1997)); divided by 0.74 m2 to get
the load capacity in kN/m2.
Load–displacement curves of mesh-reinforced shotcrete are shown in
Figure 4-18. For comparison, the results of #6-gauge weld mesh are

2
The displacements at the peak load would be less if 4’ ´ 8’ (1.2 m ´ 2.4 m) sheets were
fully bolted with 6 to 8 bolts per sheet.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 147

shown in the same figure. As discussed above, surface retaining elements


should exhibit a high initial stiffness to attract bulking loads and to
transfer them to the bolts. Mesh is relatively soft and provides little
surface support pressure at small wall displacements. This can facilitate
disruption of the stress-fractured ground near the excavation surface and
may increase the unravelling or shakedown potential. Adding shotcrete
to the mesh, i.e., forming mesh-reinforced shotcrete overcomes the issue
of low initial stiffness. It also increases the peak load carrying capacity
of the retention element. The mesh retains some of its post-peak load
capacity and mesh-reinforced shotcrete can be expected to retain a
comparable load capacity to central deflections of about 200 mm. Mesh-
reinforced shotcrete combines stiffness and ductility and thus was
labelled as “supermesh” by Kaiser et al. (1996).
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Figure 4-18 Load–displacement curves of mesh-reinforced shotcrete under static


loading (modified from Tannant et al. (1997)); divided by 0.74 m2 to get load
capacity kN/m2.
Figure 4-19 presents load–displacement curves of mesh- and fibre-
reinforced shotcrete panels for point loads, creating strain localizations
and a few cracks, and for uniform loading, creating distributed crack
patterns without localization (Kirsten and Labrum 1990). The uniaxial
compressive strengths of the shotcrete ranged from 47 to 84 MPa (55
MPa on average). The panels were bolted by four rockbolts at 1 m
spacing for a loading area of 1 m2. Fibre-reinforced shotcrete only
performs well when deformations are relatively small and the shotcrete is
uniformly strained. The peak load capacities of fibre-reinforced shotcrete
panels are smaller than those of the mesh-reinforced shotcrete. Mesh-
reinforced shotcrete panel performs better than fiber-reinforced panels,
particularly at relatively large central deflections.
Point loading is more representative of in-situ conditions than distributed
loading due to localized rock deformation. Distributed load tests create
unrealistically uniform stress and strain distributions in the test panels

Draft manuscript – Copyright protected – Cai and Kaiser 2018


148 Rock support capacity

because the load capacity of uniformly distributed loading is higher3. For


the reasons stated above, it is not recommended to use plain or fibre-
reinforced shotcrete in burst-prone ground.

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
200
100 mm thick fibre-reinforced panel,
180 uniform distributed loading
100 mm thick fibre-reinforced panel,
160 point loading
100 mm thick mesh-reinforced panel,
140 uniform distributed loading
100 mm thick mesh-reinforced panel,
120 point loading

Load (kN) 100

80

60

40

20

0
0 20 40 60 80 100 120 140 160
Deflection (mm)

Figure 4-19 Load–displacement curves of mesh- and fibre-reinforced shotcrete


panels under uniformly distributed and point loading (modified from Kirsten and
Labrum 1990); divide by 1 m2 to get load capacity kN/m2.

4.4.2 Dynamic tests


The dynamic energy absorption capacities of mesh, mesh-reinforced
shotcrete and other surface support components were investigated by
Kaiser et al. (1996) using the GRC drop test facilities. Test results on #6-
gauge weld mesh and mesh-reinforced shotcrete panels are shown in
Figure 4-20. Mesh is able to absorb about 13.5 kJ/m2 kinetic energy but
only at undesirably large central deflections. Within a desired range of
deflection of < 300 mm, mesh dissipates less than half or < 5.4 kJ/m2
kinetic energy. Mesh-reinforced shotcrete can absorb at the same
displacement threshold approximately 13.5 to 20.3 kJ/m2 of kinetic
energy (with moderate damage in the form of shotcrete cracking and
spalling at the upper end of this range).

3
It is also the reason that fibre-shotcrete promoting companies suggest the use of circular
test panels. Such tests create the most favorable but not realistic straining patterns.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 149

25
#6 mesh
severe (thin)
20 severe (thick)
moderate
minor
Kinetic energy (kJ)
15 pull test
pull test

10 mesh-reinforced
shotcrete

#6 gauge mesh

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Deflection (m)

(a)
35
#6 mesh
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

30 severe (thin)
severe (thick)
moderate
25
Kinetic energy (kJ/m2)

minor
pull test
20
pull test

15
mesh-reinforced
shotcrete
10

5
#6 gauge mesh

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Deflection (m)

(b)
Figure 4-20 (a) Impact energy versus deflection of mesh-reinforced shotcrete
and comparison to static test results (Kaiser et al. 1996); (b) unit energy capacity
(kJ/m2) obtained by dividing the bolting pattern area of 0.74 m2.
South African test data on surface support components are shown in
Figure 4-21. For design purposes, a unit energy capacity (kJ/m2) of the
surface support element is needed. The raw data given in Ortlepp and
Stacey (1998) are in kJ, not kJ/m2. Their test setups in the drop tests had
a bolt pattern of 1 m ´ 1 m. Hence, their test data of the kinetic energy
were normalized by 1 m2. For comparison, the GRC tests results for #6-
gauge mesh and mesh-reinforced shotcrete shown in Figure 4-20b are
overlayed in Figure 4-21. As mentioned above, in the GRC tests, the
impact load was applied directly as a point load in the centre of the
surface support component, whereas in the South African tests, it was
applied more uniformly by a transfer of load through concrete blocks to
Draft manuscript – Copyright protected – Cai and Kaiser 2018
150 Rock support capacity

the surface support component. In addition, there were yielding rockbolts


which absorbed some impact energy. Hence, a large difference in
deflection (at a given energy level) should be expected between the two

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
test methods and this can be observed from the data. The difference is
larger for mesh but less for mesh-reinforced shotcrete because the latter
is comparatively stiffer and can spread a load to a wider area than the
point loaded mesh. In practice, the two loading modes may be
encountered together; an individual block may be ejected and cause point
loading (i.e., GRC results are applicable) or a more or less homogeneous
package of stress-fractured ground could impact the retaining system (i.e.,
the South African test results are applicable). The practically relevant
data range for a central deflection is less than 200 mm.
When evenly loaded in a way that prevents crack localization (which is
rarely the case in practice), un-reinforced shotcrete might dissipate up to
6 kJ/m2 of kinetic energy at central deflection of < 50 mm and fiber-
reinforced shotcrete may be able to absorb energy between 15 and
20 kJ/m2 at central deflections of < 100 mm.
60
weld mesh
diamond mesh
diamond mesh and lace
50
weld mesh and lace
unreinforced shotcrete
S.F. reinforced shotcrete
40
GRC (mesh-reinforced sc)
Kinetic energy (kJ/m2)

GRC (#6 mesh)

30

20

10

0
0 100 200 300 400 500
Deflection (mm)

Figure 4-21 Energy-deflection relations of various surface support elements and


testing methods (modified from Ortlepp and Stacey (1998)). The energy in the
plot is the total energy for the support system, which include surface support
element, rockbolts, and bricks (to simulate rock mass) for the South African tests.
Data with deflection > 200 mm indicate that system incompatibilities may limit
the practical applicability.
In general, if the total wall deformation is limited to 300 mm and the bolt
deformation is restricted to 100 to 150 mm, the relative central
deformation of a surface support element will be about 150 to 200 mm.
In this deformation range, the maximum energy capacity of a surface
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 151

support is about 20 kJ/m2 (with tight, effective lacing, or mesh-reinforced


shotcrete) but typical values are less than 10 kJ/m2 for most surface
support systems with (ordinary strength) mesh.
Figure 4-22 presents the test results of mesh with lacing or mesh with ‘1-
way’ straps obtained by Stacey and Ortlepp (2001) and Ortlepp and
Swart (2002). The benefit of adding lacing or mesh straps is evident.
However, it must be realized that the reported energy values are
composed of energy dissipated by the mesh/straps and the bolts. The
stronger the mesh and the more effective the load transfer is, the more
energy will have to be dissipated by the bolts. The benefit of straps can
therefore only be gained if the bolts are strong and deformable enough to
assume the imposed load, displacement and energy demands.
The capacity of the mesh is roughly double when lacing is added.
Because of related labour cost, this is not likely cost-effective and much
heavier mesh (e.g., #4 gauge) may provide a comparable effect.
The capacity of the ‘mesh’ is roughly four times higher when 1-way
mesh straps are added. By using #0-gauge mesh straps, strength and
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

stiffness is added to the surface support system. Strength is promoted by


preventing plates from damaging wire mesh and by reducing mesh sag
between bolts. In this sense, adding straps has a similar effect as
tightening the bolting pattern4 . Reducing the sag between bolts and
consequently the central deflection of the mesh will reduce the bulking
of stress-fractured rock and restrict the unravelling process. As a
consequence, loads will be transferred more directly to the bolts and let
the (yielding) bolts absorb the remaining kinetic energy.
According to Figure 4-22, at 200 mm of central deflection, a support
system with bolted stiff mesh/strap can absorb about 50 kJ/m2 input
energies, as long as the mesh is held in place by bolts with sufficient load,
displacement, and energy dissipation capacities. Hence, when combined
with yielding bolts, mesh-reinforced shotcrete or mesh/strap
combinations can dissipate substantial dynamic energy during rock
fracturing.
A compilation of drop test data on containment support components can
be found in Potvin et al. (2010). The load, displacement, and energy
absorption capacities of surface support are summarized in Tables D-1
and D-2 in Appendix D.

4
The bolt spacing would have to be roughly halved to achieve an increase in the surface
support capacity by a factor of four.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


152 Rock support capacity

130
120 Mesh + mesh strap (Brunswick #2)

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
110 Mesh + mesh strap (Brunswick #1)
100 Mesh + mesh strap (Durastrap)

90 Special mesh and yielding rope lacing


Weld mesh and rope lacing
80
Energy (kJ/m2)

Diamond mesh and rope lacing


70
Special mesh
60
Weld mesh
50
Diamond mesh
40
30
20
10
0
0 100 200 300 400 500
Deflection (mm)

Figure 4-22 Comparison of enhancement of mesh performance by lacing and


mesh strap. The performance achieved through the combination use of yielding
bolts (1 m ´ 1 m pattern) such as conebolts tested using the South Africa facility.

4.5 Summary of rock support component


capacities
4.5.1 Reinforcement and holding components
A summary of published rockbolt support capacities is presented in
Appendix C of this volume. An effort was made to collect and
summarize data from all currently available bolt types independent of
their practical use and data quality controls. Missing data is left blank.
Data information sheets of various rockbolts can be found in Appendix E.
When choosing capacity values for design purposes from test results
(including from the table in Appendix C), it is extremely important to
account for the influence of the testing method, the energy absorption
mechanism, material properties, and shear loading as discussed earlier
and below.
Before presenting suggested properties for support design purposes,
some of the above listed factors are discussed in more detail.
Test method and facility
If test data were obtained from direct pull-out tests, they are not
necessarily representative of the capacities of rockbolts when loaded by
deep-seated rock deformations. Furthermore, it is often difficult to

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 153

compare test results because test data were obtained from various test
facilities and methods.
As discussed above, load, displacement, and energy capacities obtained
from direct impact tests differ from those obtained from indirect split-
tube tests. The difference may be substantial but each dataset has its
place in the design process. Direct pull test data are more representative
of the performance near the excavation surface due to loading via the
retention system whereas indirect tests data are more representative of
the capacity at some distances from the wall, i.e., at locations where rock
mass bulking loads the bolts. In reality, it is likely that both mechanisms
are activated to some degree but it is rather difficult to anticipate the
relative contribution to energy dissipation. For this reason, designers
often make the conservative assumption that the direct loading capacity
is all that can be relied upon (see section on support system capacity in
Chapter 5).
For example, rebar is considered as a stiff bolt with relatively low energy
absorption capacity. This interpretation is supported by numerous rebar
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

failures in support systems with wire mesh, and therefore the assumption
is correct when load is applied directly to the plate. However, tests on 22
mm rebar using the split-tube method, conducted at CANMET, show that
with a stretch length of 500 mm, the bolt could absorb 12 to 14 kJ of
energy. This is the energy capacity that is available when two rock
blocks are pulled apart inside the burst volume with 250 mm of yield or
stretch length on each side.
The free stretch or deformation zone length in fully grouted rebar largely
depends on the grout quality. Weaker grout facilitates creating longer
deformation zones and thus provides more internal energy dissipation
capacity. Weak grout, of course, is undesirable from an overall load
capacity perspective. Nevertheless, the discussion presented above
indicates that the internal energy dissipation capacity of fully grouted
rebar can be enhanced by adding stiff and strong grout (cartridges) as end
anchors for rebar with relatively weak grout along the bolt shaft (this
concept is difficult to implement in practice unless prefabricated resins
can be used for this). Other means of increasing the stretch length, e.g.,
by debonding part of the bolt length, may be more practical. Of course,
this is what the debonded threadbar and the D-bolt are aiming at.
Furthermore, the cumulative energy amounts obtained from multiple
loadings on the same bolt may differ from those of single high energy
impact loading. Multiple loadings on the same bolt are applicable to
situations where the support is repeatedly loaded by dynamic
displacement increments whereas the single high energy impact loading
is more representative for single large energy release events.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


154 Rock support capacity

Energy absorption mechanisms


Energy absorption of a bolt can be based on steel stretch (plastic
deformation of the steel), slip mechanisms (sliding device, friction) or a

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
combination of the two.
The steel stretch energy absorption capacity depends on the tensile
strength and the ultimate yield limit of the steel. In practice, it is often
limited to lower values if the bolt’s thread capacity is inferior or if the
bolt head experiences tension and bending or tension and shear. The
energy absorption ‘rate’ of a bolt, in terms of kJ/mm, is high when steel
stretch is involved but the displacement to failure and the corresponding
energy absorption depend on the free stretch length and the ultimate
strain limit of the steel. Once stretched to the limit, typically in the 50 to
200 mm/m range, a bolt will have no remnant energy absorption capacity.
Furthermore, once the steel is at the yield stage, it offers little shear
resistance.
Bolts with a slip device overcome these potential deficiencies. As long as
they are loaded in tension, they offer a much higher ultimate
displacement capacity. Their energy absorption ‘rate’ however is
generally lower. For an operationally acceptable wall displacement in the
range of 200 to 300 mm, these bolts with sliding mechanisms will have a
substantial remnant energy dissipation capacity and the remnant safety
margin may be much higher after a rockburst than when bolts relying on
steel stretch are used.
Material properties
High strength steel is needed to achieve high load capacities and both
high strength and deformability is needed to obtain high energy
capacities. High tensile strength steel exhibits high load capacity but
typically exhibit lower steel elongation capacities. Consequently, the
anticipated energy dissipation capacity of high strength steel may be
compromised, particularly when strain localization is encountered.
Steel diameter affects the axial and the shear load capacities of the bolt.
This is of particular importance when shear loading or shear localization
occurs (see Section 4.2.2).
Grout strength affects the strength and displacement capacities as well as
the stretch length of fully grouted bolts. For yielding bolts that rely on
grout-bolt friction to absorb energy, recommended grout materials must
be used and quality control is essential to consistently achieve the desired
capacities.

4.5.2 Surface retaining components


Some capacity values are presented for mesh and shotcrete (without or
with lacing or strap reinforcement), respectively, in Tables D-1 and D-2
in Appendix D. Data information sheets of surface retaining elements
can be found in Appendix E. Unfortunately, as mentioned in Chapter 5, it
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 155

is still a challenging task to test the true capacities of surface support


elements. It is also important to consider the following aspects when
utilizing the quoted capacity values for design.
Test method and facility
The energy capacities obtained from the GRC facility are for direct load
impact to the surface support element, e.g., for individual blocks being
ejected between bolts. On the other hand, the South African test setup
avoided direct contact of the drop weight with the surface support
element. Hence, an unknown part of the input energy was dissipated in
the brick pyramid. In addition, the load was spread evenly to the support
elements, creating a favourable loading condition. It is unlikely that such
uniform loading can be maintained at large displacements. The quoted
energy values from the South African test facility therefore are not the
energy capacities of the tested surface support element; they are the
system capacities. Similarly, input energy values quoted from tests by the
Swiss system are not pure retention system capacities.
Boundary condition
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

The performance of surface support elements depends on wire gauge size,


configuration, boundary condition, bolt spacing, and plate condition.
Hence, the test configuration and boundary condition can affect the result.
For example, in the South African test setup (as in the Swiss setup),
yielding bolts were used to hang the surface support element and the
brick pyramid. This creates a favourable loading condition for the
surface support elements because the yielding bolts reduced the
deflection gradient at the support elements and diminished the potential
for premature bolt-mesh connection failure. In comparison, the surface
support element was supported by stiff concrete posts in the GRC test
setup, resulting in a less favourable but often realistic loading condition
for the surface support elements in a stiff environment (e.g., when
supported by rebar).
Component interaction
It must be noted that tests conducted to obtain component capacity
values do not consider interactions between different support elements
and the interaction between rock support and rock mass. Therefore, these
capacity values must be used with caution and potential capacity
reducing effects need to be assessed. Most importantly, it is necessary to
assess the capacity at specified displacement thresholds because in most
situations the peak capacities are not simultaneously reached (see
Chapter 5).
Limitations in support capacity
Potvin et al. (2010) suggested that as a conservative design approach, the
capacity of the support system could be approximated by the capacity of
the surface support assessed using drop test. For this to be valid,
momentum transfer from broken rock would have to be the primary
Draft manuscript – Copyright protected – Cai and Kaiser 2018
156 Rock support capacity

loading process and bolts would have to be exclusively loaded by the


loads transferred from the retention system. The amount of energy
absorbed by the reinforcement and holding elements is assumed to be

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
negligible. Furthermore, it is assumed that the excavation is vulnerable
with no remnant safety margin and that the reinforced rock mass does not
contribute as an energy sink. Based on their assessment, the capacity of
the surface support is, in general, less than 10 kJ/m2. The implication of
this conservative design approach is that most designed rock support
systems would have less than 10 kJ/m2 energy dissipation capacity. As is
discussed later, this can be largely attributed to insufficient puncture load
capacities.
Their interpretation is valid if the connection between the surface support
and the holding element is the weakest link in the support system, as
fractured rock will unravel between tendons once the connection fails.
To prevent premature failure, the tendons have to be strong enough to
sustain the loads imposed by the surface elements and these loads depend
on the energy dissipation capacity of the surface support and the
stiffness/yield capacity of the bolt. For example, for a surface support
that is designed to dissipate 10 kJ/m2 and is held by one perfectly plastic
bolt per m2, the load capacity has to exceed 333 kN to maintain energy
equilibrium if the bolt has a displacement capacity of 30 mm (e.g., for a
rebar). If the displacement capacity measures 100 mm, the load demand
would have to be 100 kN. Hence, if a Split set with a load capacity of 80
kN is used, it would displace approximately 125 mm before reaching a
state of equilibrium. Even if the connections are superior, a support
system with rebar would fail and Split sets would deform excessively
(125 mm plus surface support deflection) in such a case.
In this case, the capacity of the surface system is roughly equal to the
support system capacity and therefore the extrapolation by Potvin et al.
(2010) is valid. If a support system has to dissipate more energy, the
energy dissipation capacity at the connection (bolt head) and the surface
support have to be compatible. If more energy is to be dissipated, the bolt
has to be able to dissipate energy in other ways than by load transfer
from the surface component.
Test data show that the role of a surface support element is to help/assist
the integrated support system to dissipate the total anticipated energy; it
does not have to have the same energy capacity as the rockbolt but the
rockbolt or the reinforced rock mass has to have a higher capacity than
the surface support. As discussed at various locations in this book, the
role of the retention system is to prevent unraveling between bolts.
An approach to establish the support system capacity is presented in
Chapter 5, and it is demonstrated with supporting case histories that the
capacity of a burst-resistant support system can and should be much
higher than 10 kJ/m2.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 157

4.6 Suggested design capacities for support design


Even though more recent test results are available, the reader is referred
to the CRBSHB (Chapter 4) as most of the recommendations for design
capacities are confirmed and still valid.
Here, we refine the design parameter selection to account for allowable
displacement thresholds. The ultimate capacities quoted in the CRBSHB
can rarely be reached and it is therefore more meaningful to design for a
chosen displacement threshold. This is particularly relevant when
support components are combined into a support system where not all
support components are equally loaded and deformed (see Chapter 5).
For this reason, energy values are listed for 100 and 200 mm
deformations. The energy values for tendons represent the maximum
available capacities, i.e., they are only valid if the tendons have not been
preloaded. Furthermore, they can only be mobilized if the puncture load
capacity of the surface support is higher than the load needed to yield the
tendons or if the tendons are directly loaded by impacting rock.
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

4.6.1 Design capacities of reinforcing and holding components


in burst-prone ground
Because many influence factors are not consistently treated in support
element testing, inconsistencies in capacity values quoted in Table C-1
(obtained from literature) need to be adjusted based on field observations
and qualitative assessments to arrive at capacity values for design
purposes. For the following, Sections 4.6.1 and 4.6.2, the authors used
their engineering judgment to establish representative values for rock
support components that can be adopted for support system designs.
From the discussion presented in this chapter, it is evident that the
remnant energy capacity of support components depends on the
previously imposed displacements. As will be further elaborated in
Chapter 5, it is not the ultimate or maximum capacity but the capacity of
an individual component at a practically relevant displacement threshold
that matters and needs to be considered in a design. As a matter of fact,
as not all components in a support system are simultaneously loaded to
their capacities, it is meaningless to use the ultimate load and energy
capacities in a design; it is also meaningless to use the combined capacity
(by simple addition) to obtain the system’s capacity. For this reason, a
deformation-based approach is advocated in Volume II, and it is
recommended to select compatible support components for specific
displacement thresholds – thresholds that are smaller than the ultimate
displacement capacities.
Table 4-2 lists the direct loading (loading at plates) capacities (peak and
average load capacities) of some widely used rockbolts and their energy
capacities for two displacement ranges: 0 to 100 and 0 to 200 mm,
respectively; they are called E100 and E200.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


158 Rock support capacity

For the purpose of establishing E100 and E200, the load displacement
curves are approximated by an equivalent perfectly plastic model as
shown by example in Figure 4-23 for three bolt types (cemented rebar,

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
D-bolt, and Super Swellex). The area under the red line that indicates the
energy capacity of the bolt is equal to the area under the actual load
displacement curve, which represents the energy capacity of the bolt. The
‘average load’ therefore signifies the load that ‘on average’ reflects the
load capacity for a given bolt and a chosen displacement range. For
example:
- For the cemented rebar with a peak capacity of 180 kN, the
‘average load’ measures 170 kN over a displacement range of 35
mm and the corresponding energy capacity of E = 6 kJ.
- For the Super Swellex with a peak capacity of 120 kN, the
‘average load’ measures 95 kN over a displacement range of 0 to
140 mm and the corresponding energy capacity of E = 13 kJ.
Assuming that the average load is maintained to 150 mm, the
energy capacity of this bolt is E = 14.3 kJ.
- For the D-bolt with a 1 m deformation section and a peak
capacity of 220 kN, the ‘average load’ amounts to 190 kN over a
displacement range of 10 to 160 mm and the corresponding
energy capacity of E = 28.5 kJ. For a deformable section
between paddles or between plate and first paddle of 0.5 m, the
energy capacity would be about half or 14.3 kJ due to the
reduced displacement capacity.

Figure 4-23 Approximations of load–displacement relation by perfectly plastic


models (red capacity lines) for indirect loading examples (modified Figure 15
from Li (2010)).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 159
Table 4-2 Direct loading capacities of tendons for burst-resistant support
design
Bolting type Peak [‘average’ E100 E200 Emax dmax Note
load] capacity (kN) (kJ) (kJ) (kJ) (mm)
Mechanical bolt 140 [70] 1.5–3 N/A 1.5–3 20–40 Capacity highly
17.3 mm variable depending on
anchorage; not
recommended for
burst-prone ground

Rebar 22 mm 230 [175] 1.5–3.5 N/A 1.5–3.5 10–20; If perfectly grouted to


(cement grouted) collar.
230 [200] 3–8 3–8 15–40 With partial grouting
near collar or stress
relaxation
Threadbar
19.5 mm; non- 180 [150] 1.5–4.5 N/A 1.5–4.5 10–30 If perfectly grouted to
debonded collar.
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

1.6 m debonded 180 [165] 10–15 N/A 10–15 60–90 With partial grouting
near collar or stress
relaxation
Smooth bar 16 mm 120 [85] 5–9 N/A 5–9 60–110 Fully grouted
fully grouted

Cablebolt 12.7 mm 185 [165] 3–6 N/A 3–6 20–35 Fully grouted
15.2 mm 260 [235] 5–9.5 N/A 5–9 20–40

Cablebolt 15.2 mm
- debonded (4 m) 200-250 [195] 20–25 N/A 30–35 150–175
- dynamic 80-180 [100] 10 20 30 300
- duracable 70-100 [85] 7–10 14–20 > 19 > 220
Split Set 39 mm 20-30 [25] 2–3 4–5 2–8 100–250 per 1 m embedment
46 mm 25-40 [32] 3–4 5–8 5–10 150–300 length

Swellex Mn 12 50-70 [60] 5–7 10–14 6–14 100–200 per 1 m embedment


yielding Mn 24 80-120 [100] 8–12 15–25 10–35 100–300 length
Conebolt 16 mm 100 10 20 20 > 200 Fully grouted
22 mm 200 20 40 40 > 200
Modified conebolt 130 [115] 11–13 23–26 > 30 250–500 Fully grouted
17.3 mm
D-bolt 20 mm 220 [190] 10–15 N/A 10–15 50–75 If direct loaded at plate
22 mm 260 [220] 11–17 N/A 11–17 50–75 with < 0.5 m stretch
length; 10–15% yield
capacity at thread

Note:

Draft manuscript – Copyright protected – Cai and Kaiser 2018


160 Rock support capacity

- This table is applicable for bolts loaded by the surface support or


by rock impacting the plate.
- The load that produces Emax at the dmax is the ‘average’ load

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
capacity [ ]. This load value is used later to establish the capacity
of integrated support systems.
- If the bolt spacing is 1 m ´ 1 m, the quoted values are per m2;
otherwise adjustments have to be made to obtain the capacity per
m2 .
For bolts with ultimate displacement capacities less than 100 mm, E100 is
equal to the ultimate energy capacity Eult because it represents the energy
consumed before the displacement threshold is reached. Even if a bolt
fails at or before the 100-mm displacement threshold, it still dissipates
energy and thus contributes to the energy dissipation capacity of the
integrated support system. This is of particular importance when bolts
with low displacement capacities are installed during rehabilitation or
proactive support maintenance work. Because they have not yet been
strained by mining-induced displacements, they offer their ultimate
energy dissipation capacity during future burst loading.
These tables are based on the authors’ interpretations of available test
data and contain recommendations for design parameters to establish the
capacity of integrated, burst-resistant support systems. They are intended
for a support system design at the prefeasibility stage and will be used in
Chapter 5 for support system capacity calculations.

Table 4-3 presents the corresponding capacities for indirect loading. Not
all rockbolts work in the same manner during indirect loading inside the
bursting rock mass. For example, conebolts are always loaded at the
plate and cone anchor, cables or rebar are stretched (pulled out of rock)
on both sides of the internally separating rock blocks (at stress-fractures),
and friction bolts are pulled-out as in direct loading.
As indicated in Section 4.2.2, it can be concluded from static shear tests
that the dowelling capacity is less than about 2/3 of the pure tensile
capacity. For rockburst support design purposes, it is necessary to
consider that the displacement capacity and thus the energy capacity may
be drastically reduced when shear localization occurs. The practical
displacement and energy capacities may be much lower than the values
listed in the above presented tables.
There are many other factors that affect the quoted design values and it is
therefore prudent to evaluate a design resulting from these parameters
with experience at a given mine. Furthermore, it is essential that a design
is verified by in situ monitoring and revised once field evidence suggests
that adjustments are required.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 161
Table 4-3 Indirect loading capacities of tendons for burst-resistant support
design
Bolting type Peak [‘average’ E100 E200 Emax dmax
load] capacity (kJ) (kJ) (kJ) (mm)
(kN)
Rebar 22 mm 230 [190] 4–8 N/A 4–8 20–40 Perfectly grouted to collar;
6–15 N/A 6–15 30–80 partial grouting or stress
relaxation
Threadbar 19.5 mm 180 [170] 4–10 N/A 4–10 20–60 Fully grouted
Smooth bar 16 mm 120 [85] 7–9 7-17 7–20 80–250 Fully grouted
fully grouted
Cablebolt 12.7 mm 185 [160] 6–11 N/A 6–11 40–70
15.2 mm 260 [230] 9–16 N/A 9–16 40–70

Spilt Set 39mm 20-30 [25] 2.5 5 2–8 100–250 per 1 m length
46 mm 25-40 [32] 3.2 6.4 5–10 150–300
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Swellex Mn 12 50-70 [60] 5–7 10–14 6–14 100–200 per 1 m length


yielding Mn 24 100-120 [120] 8–12 15–25 10–35 100–300

Conebolt 16 mm 100 10 20 20 > 200 Fully grouted


22 mm 200 20 40 40 > 200
Modified conebolt 130 [115] 11–13 23–26 > 30 250–500 Fully grouted
17.3 mm

D-bolt 20 mm 220 [190] 19 29–38/m 29–38/m 150–200 for 1 m stretch length


22 mm 260 [220] 22 33–44/m 33–44/m per m between paddles; 15-20%
yield capacity

4.6.2 Design capacities of retention components


The capacity of a surface support system strongly depends on the bolt
patterns (spacing). For this reason, the range of design parameters
provided in Table 4-4 covers practically meaningful patterns from
relatively wide (approximately 1.5 m2/bolt) to relatively tight (approx.
0.75 m2/bolt). The listed design values are intended for preliminary
designs and should be confirmed by field observations and back-analyses
of support damaged by rockbursts.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


162 Rock support capacity
Table 4-4 Retention system capacities for burst-resistant support design
Retention system type Puncture E100 E200 Emax
(range for wide (> 1 m2) and narrow (< 1 load (kN) (kJ/m2) (kJ/m2) (kJ/m2)

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
m2) bolting pattern) at maximum
deflection

Light to moderate weld mesh 10–25 0.5–1 3–5 5–10


Moderate to heavy weld mesh 25–40 1–2 5–8 10–15
#9 Chain-link mesh 20–40 ≤1 < 1–2 10–15
Moderate mesh with heavy mesh 25–40 3–6 10–15 20–40
straps
Mesh-reinforced shotcrete (nominal 15–40 5–10 10–15 20–30
thickness ≤ 100 mm)
with #0 mesh straps (or heavy mesh) 40–80 10–15 15–25 25–40
Fibre-reinforced shotcrete (nominal 5–10 ≤ 1–2 1–3 < 2–5
thickness ≤ 100 mm) 5

Note:
- E100 and E200 are energy values at 100 and 200 mm central
deflections, respectively. The energy capacity measures less than
the listed value if the bolt spacing is greater than 1.2 m.
- This table is applicable if the weakest link issue is eliminated
and tendons provide at least the listed puncture load.
- The puncture load is essentially independent of bolt spacing but
the quoted values are typically reached at displacements
exceeding 200 mm.
Because most bolt types will not yield at these puncture loads, the values
listed in this table constitute the support system capacity when broken
rock exclusively impacts the surface support. The system capacity is
higher when bolt plates are directly impacted. Retention system data
reported by Potvin et al. (2010) suggest that puncture loads rarely exceed
20 kN and corresponding energy values rarely exceed 10 kJ/m2. This is
to be expected when broken rock loads the support between the tendons.
Table 4-4 indicates that, for an operational range of 100 to 200 mm
central deflection, standard strength mesh alone provides inadequate
burst-resistance at < 10 kJ/m2 energy capacity. Hence, mesh with heavy
straps, high strength mesh or mesh-reinforced shotcrete must be

5
Tests with the South African system producing distributed loading (with distributed
shotcrete crack patterns) and damping of impact energy by breaking concrete blocks
suggests values as high as 10 to 15 kJ/m2 at central deflections of 100 to 200 mm.
Because these values contain other than surface system energy dissipation components,
they are not considered to be representative of pure retention system capacity values.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 163

considered as a minimum surface support standard for burst-prone


ground. If superior retention is desired, straps or mesh over mesh-
reinforced shotcrete are needed. These will roughly double the retention
capacity with respect to puncture load and energy dissipation. Within the
desirable deformation range of 100 to 200 mm central deflection, an
energy capacity of about 20 kJ/m2 can then be relied upon. For burst-
prone ground, fibre-reinforced shotcrete provides insufficient puncture
load and energy capacities and mesh over or mesh and straps over fibre-
reinforced shotcrete must be used.

4.6.3 Design capacity compatibility


How to use the capacity tables to arrive at the capacity of integrated
support systems is explained in more detail in Chapter 5. For the simplest
case of direct loading, when fractured rock impacts the support via the
retention system, the energy capacities of the retention and holding
components can be assumed to work in parallel, but only if all the
tendons reach the yield stage, i.e., at the ‘average’ load listed in Table
4-2. Hence, the values listed in Table 4-2 and Table 4-4 can be combined
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

(added) if the bolt forces are sufficient to yield the tendons.


Unfortunately, this is often not the case as the puncture loads are
frequently much less than the ‘average’ bolt capacities (compare the
loads in [ ] in Table 4-2 and the puncture loads in Table 4-4).
In other words, a key design capacity compatibility criterion is that the
puncture load capacity must exceed peak load capacity of the individual
tendons.
It follows that, unless bolts are directly impacted by bursting rock, most
bolt types will not yield before the puncture load is reached. The
retention system data reported by Potvin et al. (2010) confirms this,
suggesting that puncture loads rarely exceed 20 kN. As a consequence,
the corresponding energy values rarely exceed 10 kJ/m2.
This is to be expected when broken rock loads the support between
tendons. For example, any support system with a moderate to heavy weld
mesh would only be able to dissipate E200 (mesh) = 5 to 8 kJ/m2. If
suspended by 39 mm Split sets that could yield or slip before the
puncture load is reached, the energy capacity at 300 mm (for a 1 m ´ 1 m
pattern with at least 1 m of anchor length) would be slightly higher at
E300 (mesh plus Split sets) = 7 to 11 kJ/m2 (8 to 12 kJ/m2 with 46 mm
Split sets).
These rather limited energy capacities can only be exceeded if the
puncture load is substantially higher than the 20 kN value quoted by
Potvin et al. (2010) such that the bolts yield and dissipate energy. Table
4-2 shows that puncture load capacities must exceed the following values
to ensure energy dissipation by the bolting system:
- 30 kN/m anchor length for Split sets;

Draft manuscript – Copyright protected – Cai and Kaiser 2018


164 Rock support capacity

- 60 kN/m anchor length for Swellex bolts;


- 85 kN for smooth bars;

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
- 115 kN for conebolts; and
- more than 150 kN for rebar, threadbar, most cablebolts, and D-
bolts.
According to Table 4-4, none of the listed retention systems has
sufficient puncture load capacity other than for Split sets with relatively
short anchor length.
In other words, a high puncture capacity is the key to achieving a high
energy dissipation capacity and thus a successful burst-resistant support
system. This can only be achieved by combining some of the retention
components listed in Table 4-4 and by adding large plates. For example,
by adding moderate mesh with heavy mesh straps with large plates to
mesh-reinforced shotcrete (nominal t ≤ 100 mm), the puncture capacity
can be increased to 80 to 120 kN or more if the plate is large enough to
prevent puncture failure. High strength mesh may be needed to reach
puncture loads for high capacity / large diameter rebar, threadbar,
cablebolts, and D-bolts (see next on the discussion of the El Teniente
double-layer system).
When part of the support system is indirectly loaded, energy dissipates
near the wall and inside the rock. This scenario with complex support
loading mechanisms is discussed in Chapter 5.
Assessment of El Teniente double-layer system
El Teniente mine developed and now widely adopts a double-layer
support system (Figure 4-24) with high strength chain-link mesh over
mesh-reinforced shotcrete held with heavy rebar (at 1.2 m ´ 1.2 m square)
and cablebolts (at 2 m ´ 2 m square) in areas with a high burst hazard.
The high strength mesh alone, as used in the double-layer retention
system, has a static puncture capacity of 180 kN at 310 mm central
deflection with a 1.3 m ´ 1.3 m bolt pattern.
From impact tests on mesh supported by bolts, Bucher et al. (2013)
estimated the percentage of energy split between the test frame, rockbolts,
high strength mesh, and the test frame. Of the total energy input, 16%
was consumed by the yielding bolts and only 5% by the mesh (78% by
frame). In the setup using a soft concrete slab, the energy distribution
was 5% by the bolts, and 13 % by the mesh (82% by frame). This
suggests that 5 to 13% of the impact energy was dissipated by the surface
support.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 165
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

(a) (b)
Figure 4-24 (a) Double-layer surface support system adopted by El Teniente
mine and tested in the Swiss facility with the layout shown in (b) (Brändle et al.
2017).

The system shown in Figure 4-24 was impact-loaded causing centre


deflections between 240 to 600 mm (Munoz et al. 2017). During impact-
loading in a central area of 1 m2 with a drop-weight excreting 60 kJ/m2,
anchor reaction forces between 155 and 429 kN were measured (Brändle
et al. 2017) (1.2 m ´ 1.2 m pattern). Because none of the bolts or cables
punched through the double-layer system, it follows that puncture load
capacity of the double-layer system was > 180 kN and it is likely that
some of the bolts temporarily yielded. Nevertheless, considering the
damage to the retention system, it is possible that most of the impact
energy was consumed by the surface support and the test frame.
For an assumed energy split of 75% by the frame and 25% or 15 kJ/m2
by the bolts and surface support, and a 25:75% to 75:25% split between
bolts and the retention system (range reported by Bucher et al. 2013), it
follows that between 4 and 11 kJ/m2 were consumed by the double-layer
surface support (15 kJ/m2 if zero energy was consumed by the bolts).
This energy dissipation capacity for the surface support is at the lower
end of the values given in Table 4-4.
It can therefore be concluded that the double-layer system has a puncture
load capacity in excess of > 180 kN and an energy dissipation capacity of
about 10 to 15 kJ/m2.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


166 Rock support capacity
Table 4-5 El Teniente double layer retention capacities for burst-resistant support
design
Retention system Puncture load E100 E200 Emax

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
(kN) at maximum (kJ/m2) (kJ/m2) (kJ/m2)
deflection

El Teniente double-layer > 180 5–7 9–13 10–15


surface support system (estimated) (estimated)

A comparison with the values listed in Table 4-4 shows that the greatest
benefit of the double-layer system is the far superior puncture load
capacity. The energy dissipation capacity at 100 to 200 mm central
deflection is, however, comparable to mesh-reinforced shotcrete without
or with mesh straps.
Because the weak link issue has been resolved by providing a very high
puncture capacity, and the double-layer surface support alone should be
able to absorb between 10 and 15 kJ/m2. The energy dissipation capacity
of the integrated system consists of the energy capacity of the bolt
system at a given displacement threshold plus 10 to 15 kJ/m2. The
maximum direct energy dissipation capacity of the integrated double-
layer support system (Figure 4-24) with 25 mm threadbar at 1.2 m ´ 1.2
m and 15.2 mm twin cablebolts at 2 m ´ 2 m spacing is therefore in the
order of 15 to 24 kJ/m2. If the threadbar were debonded over 1.6 m and
the cables over 4 m, the maximum integrated support system capacity
would be about 31 to 45 kJ/m2. These estimates are conservative but
much more realistic than cumulative maximum capacity estimate of 88
kJ/m2 quoted by Munoz et al. (2017). The corresponding energy
dissipation estimates for 100 and 200 mm central deflections are listed in
Table 4-6.
This example also confirms that the maximum practical energy support
limit (MPESL) is 50 kJ/m2.
Table 4-6 Direct loading capacities of El Teniente double-layer system
per m2 and for 100 or 200 mm central deflections
Bolt debonding Peak [‘average’ E100 E200 Emax dmax
load] capacity (kJ/m2) (kJ/m2) (kJ/m2) (mm)
type
(kN)
Without debonding > 180 8–12 15–24 15–24 250

With debonding > 180 10–15 20–30 31–45 300


(estimates) (estimates)

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 167

In summary, it is necessary to ensure at all times that the forces


generated during energy dissipation by the surface element are
compatible with the load capacity of the holding tendons. The surface
components can only dissipate the energy listed in Table 4-4 if the
tendons are strong enough to hold the dynamically loaded surface
element. On the other hand, the tendons can only dissipate energy if the
puncture load capacity is high enough to cause yield in the tendons.
When part of the support system is directly loaded by rock, the energy
dissipation capacity is higher, and if indirectly loaded, energy is
dissipated near the wall and inside the rock. This scenario with complex
support loading is discussed in Chapter 5.

4.7 References

Ansell, A. 1999. Dynamically loaded rock reinforcement. Doctoral thesis,


Dept. of Structural Engineering, Royal Institute of Technology;
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Stockholm. p.
Ansell, A. 2005. Laboratory testing of a new type of energy absorbing rock
bolt. Tunnelling and Underground Space Technology 20(4): 291-300.
ASTM. 2008. ASTM D7401-08, Standard Test Methods for Laboratory
Determination of Rock Anchor Capacities by Pull and Drop Tests
(Withdrawn 2017), ASTM International, West Conshohocken, PA, 2008,
www.astm.org.
Aziz, N., Pratt, D., and Williams, R. 2003. Double shear testing of bolts. In
Coal Operators' Conference, University of Wollongong. pp. 154-161.
Bjurstrom, S. 1974. Shear strength of hard rock joints reinforced by grouted
untensioned bolts. In Proc. 3rd ISRM Cong. , Denver. 2 pp. 1194-1199.
Brändle, R., Rorem, E., Luis, R., and Fisher, R. 2017. Full-scale dynamic
tests of a ground support system using high-tensile strength chain-link
mesh in El Teniente mine, Chile. In 1st Internation Conf. on
Underground Mining Technology ACG. pp. 25-43.
Bucher, R., Cala, M., Zimmerman, A., Balg, C., and Roth, A. 2013. Large
scale field tests of high-tensile steel wire mesh in combination with
dynamic rockbolts subjected to rockburst loading. In Ground Support.
Edited by B.G.H. Brady and Y. Potvin. pp. 221-232.
Cai, M., and Champaigne, D. 2012. Influence of bolt-grout bonding on MCB
conebolt performance. Int. J. Rock Mech. Min. Sci. 49(1): 165-175.
Cai, M., Champaigne, D., and Kaiser, P.K. 2010. Development of a fully
debonded conebolt for rockburst support. In 5th International Seminar
on Deep and High Stress Mining. Edited by M. Van Sint Jan and Y.
Potvin, Santiago, Chile. pp. 329-342.
Cai, M., Kaiser, P.K., Suorineni, F., and Su, K. 2007. A Study on the
Dynamic Behaviour of the Meuse/Haute-Marne Argillite. Physics and
Chemistry of the Earth 32(8-14): 907-916.
Doucet, C. 2012. Ground Support Research at Canmet Mining. In WSN
Symposium on dynamic ground support applications, Sudbury, Ontario.
Draft manuscript – Copyright protected – Cai and Kaiser 2018
168 Rock support capacity
Gomez, J.T., Shukla, A., and Sharma, A. 2001. Static and dynamic behavior
of concrete and granite in tension with damage. Theoretical and Applied
Fracture Mechanics 36: 37-49.

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Kaiser, P.K., Tannant, D.D., and McCreath, D.R. 1996. Canadian Rockburst
Support Handbook. Geomechanics Research Centre, Laurentian
University, Sudbury, Ontario. p. 314.
Kirsten, H.A.D., and Labrum, P.R. 1990. The equivalence of fibre and mesh
reinforcement in the shotcrete used in tunnel-support systems. J. South
Afr. Inst. Min. Metall. 90(7): 153-171.
Li, C. 2010. A new energy-absorbing bolt for rock support in high stress
rock masses. Int. J. Rock Mech. Min. Sci. 47: 396-404.
Li, C., and Charette, F. 2010. Dynamic performance of the D-Bolt. In Proc.
5th Int. Seminar on Deep and High Stress Mining. Edited by M. Van
Sint Jan and Y. Potvin. pp. 321-328.
Malvar, L.J., and Crawford, J.E. 1998. Dynamic increase factors for steel
reinforcing bars. In 28th DDESB Seminar. Orlando, USA. pp. 1-17.
Morton, E.C., G., T.A., and Villaescusa, E. 2009. The performance of mesh,
shotcrete and membranes for surface ground support. In ROCKENG09:
Proceedings of the 3rd CANADA-US Rock Mechanics Symposium.
Paper 4002.
Munoz, A., Rojas, E., Brandel, R., Luis, R., and Fisher, G. 2017. Full-scale
dynamic tests of a ground support system using two layers of high-
tensile strength chain link mesh to increase the energy absorption at EL
Teniente Mine, Chile. In RaSiM9, Santiago, Chile. pp. 159-167.
Ortlepp, W.D., and Erasmus, P.N. 2005. Dynamic testing of a yielding cable
anchor. In 3RD Southern African Rock Engineering Symposium.
Ortlepp, W.D., and Stacey, T.R. 1998. Performance of tunnel support under
large deformation static and dynamic loading. Tunnelling and
Underground Space Technology 13(1): 15-21.
Ortlepp, W.D., and Swart, A.H. 2002. Performance of various types of
containment support under quasi-static and dynamic loading conditions,
Part II. p. 100.
Player, J.R. 2012. Dynamic Testing of Rock Reinforcement Systems.
Western Australian School of Mines, Curtin University of Technology.
PhD Thesis. p. 501.
Player, J.R., Thompson, A.G., and Villasescusa, E. 2009. Dynamic testing of
threadbar used for rock reinforcement. In ROCKENG09: Proceedings of
the 3rd CANADA-US Rock Mechanics Symposium. Edited by M.
Diederichs and G. Grasselli, Toronto. Paper 4030.
Player, J.R., Villasescusa, E., and Thompson, A.G. 2004. Dynamic testing of
rock reinforcement using the momentum transfer concept. In Ground
Support in Mining and Underground Construction. Edited by E.
Villasescusa and Y. Potvin. paper 29.
Potvin, Y., Wesseloo, J., and Heal, D. 2010. An interpretation of ground
support capacity submitted to dynamic loading. In 5th International
Seminar on Deep and High Stress Mining. Edited by M. Van Sint Jan
and Y. Potvin, Santiago, Chile. pp. 251-272.
Scott, C., Penney, A.R., and Fuller, P. 2008. Competing factors in support
selection for the west zone of the Beaconsfield Gold Mine, Tasmania. In
Narrow Vein Mining Conference, Ballarat, Vic. pp. 173-178.
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 169
Snell, G., Kuley, E., and Milne, D. 2017. A laboratory-based approach to
assess rockbolt behaviour in shear. In Underground Mining Technology
2017. ACG, Sudbury. pp. 45-54.
Stacey, T.R., and Ortlepp, W.D. 1999. Retainment support for dynamic
events in mines. In Rock Support and Reinforcement Practice in Mining.
Edited by E. Villaescusa and C.R. Windsor and A.G. Thompson. A.A.
Balkema. pp. 329-333.
Stacey, T.R., and Ortlepp, W.D. 2001. Tunnel surface support capacities of
various types of wire mesh and shotcrete under dynamic loading. J.
South Afr. Inst. Min. Metall.: 337-342.
Stillborg, B. 1994. Professional users handbook for rock bolting. 2nd ed.
Clausthal-Zellerfeld: Trans Tech Publications. p.
Tannant, D.D., Kaiser, P.K., and Maloney, S. 1997. Load-displacement
properties of welded-wire, chain-link, and expanded metal mesh. In
Proceedings of the International Symposium on Ground Support,
Norway. pp. 651-659.
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Draft manuscript – Copyright protected – Cai and Kaiser 2018


170 Rock support capacity
Rockburst Support Reference Book (I) 171

Chapter Five
Rock support system capacity
Synopsis
As discussed in previous chapters, the capacities of individual
support elements (rockbolt, mesh, or shotcrete) can be determined
using laboratory and field tests. However, most testing facilities
are not suitable to test the capacity of integrated support systems,
in particular, the interaction between rock support and rock. In
general, the effectiveness of a rock support system can only be
verified by ground truthing, i.e., by performance assessments dur-
ing dynamic loading by actual rockbursts.
Even though it is difficult to provide the design capacity of a rock

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
support system, we offer here a means to estimate the support sys-
tem capacity. A new approach to establish the capacity of a rock
support system consisting of various support components is intro-
duced and the maximum practical support limits in terms of dis-
placement and energy capacities of rock support systems are pre-
sented.
It is recognized that some of the support’s capacity is consumed as
mining-induced static or dynamic displacements accumulate.
Hence, the remnant capacity at the time of a rockburst is typically
much less than that of the installed support and it may be neces-
sary to supplement a support system’s capacity in a timely fashion
so as to prevent excavation damage. For this purpose, the concept
of proactive (or preventive) support maintenance (PSM) is intro-
duced as a potentially cost-effective means to ensure support ef-
fectiveness and thus safety in seismically active mines at the time
a rockburst occurs.

5 Rock support system capacity ................................................... 173


5.1 Rockburst damage mitigation ............................................. 173
5.2 Integrated support system characteristics ............................ 175
5.2.1 Support to mitigate rockburst damage dominated by
dynamic disturbances from remote seismicity (direct loading) ... 177
5.2.2 Support to mitigate strainburst damage dominated by
bulking and shear rupture (direct and indirect loading) .............. 180
5.2.3 Rock support system testing – simulated rockbursts .... 183

Draft manuscript – Copyright protected – Cai and Kaiser 2018


172 Rock support system capacity

5.3 Estimation of support system capacity ................................ 186


5.3.1 Support capacity consumption..................................... 187
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

5.3.2 Proactive support maintenance .................................... 189


5.3.3 Direct impact loading capacity of bolting system......... 193
5.3.4 Contribution of surface support to integrated support
system capacity ......................................................................... 202
5.3.5 Combined direct impact loading capacity of integrated
support systems ......................................................................... 207
5.3.6 Maximum practical energy support limit (MPESL) ..... 209
5.3.7 Indirect impact loading capacity of bolting systems..... 210
5.4 Rock support systems – ground-truthing ............................. 215
5.4.1 Big Bell Mine, Australia ............................................. 215
5.4.2 Brunswick Mine, Canada ............................................ 217
5.4.3 Copper Cliff North Mine, Canada ............................... 221
5.4.4 Kidd Mine, Canada ..................................................... 222
5.4.5 El Teniente TEN-SUB6, Chile .................................... 224
5.4.6 Grasberg mine, Indonesia............................................ 228
5.4.7 Comments on other case examples presented in the
CRBSHB .................................................................................. 232
5.5 References .......................................................................... 232

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 173

5 Rock support system capacity


Support systems are made up of various support components with
the characteristics described in Chapter 4. In this chapter, the sup-
port system loading process is first reviewed and available system
testing methods are discussed. Whereas the capacities of individ-
ual support elements (rockbolts, mesh, or shotcrete) can be deter-
mined using laboratory and field tests, most testing facilities are
not suitable to test the capacity of a rock support element in an in-
tegrated support system. In particular, the interaction between rock
support and rock is missing in these tests. Hence, establishing the
capacity of an integrated rock support system or a reinforced rock
mass is challenging.

5.1 Rockburst damage mitigation


Support systems are commonly designed to dissipate energy from

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
a dynamic disturbance to prevent or mitigate damage from rock-
bursts. Whereas design against rock ejection by energy transfer
from seismic events is a necessary element of support design, it is
not necessarily a sufficient criterion. There are other energy
sources (e.g., stored strain energy in the rock mass surrounding the
burst volume) and other energy sinks beyond the support’s energy
dissipation capacity (i.e., fracture and frictional energy sinks) that
need to be considered. In fact, when an excavation wall fails dur-
ing a self-initiated, triggered or dynamically loaded strainburst, it
is the tangential stress and radial deformation that bring the sup-
ported rock mass to failure as illustrated by Figure 5-1a, causing a
sudden inward displacement d of the broken rock (see also Section
3.2.3). Therefore, the support has to be designed to survive burst
related displacements and associated strain localizations. Finally,
it has to be recognized that the energy dissipation and the dis-
placement capacity is being consumed as a support system gets
deformed by mining-induced stress-fracturing and rock mass de-
formations.
A proper support system has to deform to accommodate rock mass
failure processes and provide stabilizing forces to the surrounding
rock mass both in the radial and in the tangential directions, as il-
lustrated by the blue arrows in Figure 5-1b. If a support system is
effective, it will survive the dynamic deformations and the entire
package of supported and reinforced ground will collectively dis-
sipate the released energy. With an effective support system, there
will be no ejection, i.e., the terminal ‘ejection’ velocity of the sup-
ported ground is zero. It is the role of the integrated support sys-
tem to lower ground motions to zero by dissipating energy in the
deformation process of the reinforced rock mass. In other words,

Draft manuscript – Copyright protected – Cai and Kaiser 2018


174 Rock support system capacity

the design objective is to create a deformable reinforced rock vol-


ume (called gabion) that provides the rock mass a self-support ca-
pability during the entire dynamic failure process, i.e., as the gabi-
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

on is deforming in the post-peak deformation stage.

(a) (b)
Figure 5-1 (a) Forces acting on a volume of burst-prone rock and direction of resulting
bulking displacement d; (b) deformable support system for ground control in walls of
drifts indicating resistance forces of a reinforced support system to resist convergence
(vertical arrows) and confine the surrounding rock mass (horizontal arrows) (Kaiser
2017).
For this failure process, the support system cannot be planned
based on energy demands from remote seismic events alone. A
deformable support system has to be designed to compensate en-
ergy release from the stored strain energy in the surrounding rock
mass by providing a number of energy sinks. These energy sinks
consist of energy consumed by fracturing and shearing the rein-
forced rock mass and deforming the installed support components.
In the following discussion, internal energy sinks are ignored. In-
stead, a deformation-based support design approach that deals
with the displacement-dependent-support-capacity mobilization
and consumption is presented to establish the direct loading capac-
ity of rock support systems. A deformation-based support design
approach that considers internal energy sources and sinks during
strainbursting is presented in Volume II.
Independent of the design approach, deformations have to come to
rest to achieve a new equilibrium. Therefore, the support system
has to provide a deformable gabion of stress-fractured rock that
cannot unravel, deliver sufficient confinement to the surrounding
ground, and provide resistance in the tangential direction to mini-
mize the effective excavation span (blue arrows in Figure 5-1b).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 175

5.2 Integrated support system characteristics


Even though it is difficult to provide an accurate estimate of the
design capacity of support systems, the authors offer here a means
for estimating the support system capacity by considering the cu-
mulative shared contribution of each support element. This chap-
ter also introduces ways to estimate how the support capacity is
consumed by mining-induced displacements and methods to re-
store support system capacity by proactive support maintenance
(PSM).
During a rockburst, energy is absorbed by the entire system (Esys-
tem), not just by a single support element. Part of the input energy
is absorbed by breaking and deforming the reinforced rock mass
(Erock mass), part by deforming the surface support (Esurface) and the
bolts at the plate (Ebolt plate), and part by straining the reinforcement
inside the rock mass (Ereinforcement ).
Accordingly, the total reinforced rock mass system capacity con-

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
sists of:
Esystem = SEbolt plate + Esurface + Erock mass + SEreinforcement (5-1)

where each energy term is a function of the imposed displacement


at the time the system energy capacity is determined. The system
energy capacity is therefore not constant and depends on the range
of support components used and the installation sequence. The
summation symbol S indicates that the rockbolt term is typically
composed of energy capacities of various bolt types.
The percentage split between these four energy sinks is largely
unknown and it depends on the dynamic loading process as illus-
trated by Figure 5-2.
In an effective support system, where failure is prevented as long
as the connections to the bolting system are of superior capacity, it
is reasonable to assume that Esystem >> (SEbolt plate + Esurface) because
Erock mass and Ereinforcement are not zero. Eq. (5-1) can be rewritten as:
Esystem = S(Ebolt plate + Ereinforcement) + Esurface + Erock mass (5-2)
where, Ebolt = S(Ebolt plate + Ereinforcement) =aB Esystem;
Esurface = aS Esystem; and
Erock mass = aR Esystem, such that aB + aS + aR = 1.
When a reinforced rock mass is loaded by tangential straining,
much energy is consumed by overcoming the cohesive strength
and friction and aR > 0. For ultra-conservative designs, aR is often
set to zero as the energy absorbed by breaking the rock mass is
ignored. In strainbursting ground, it is certainly reasonable to as-
sume aR = 0 or a R equalling a relatively small value because the

Draft manuscript – Copyright protected – Cai and Kaiser 2018


176 Rock support system capacity

failing rock is part of the source of energy; hence, at least for


aB + aS ≈ 1.
strainbursting ground,
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

ent
ote ev
om rem Stra
d fr inbu
ic loa rst 1
Dynam

3 VSB
1

Str
a
inb
urs
V SB2

t 2


SB1 … ejects wedge of broken rock and directly hits bolts
SB2 … indirectly loads (strains) bolts and ejects wedges

Figure 5-2 Dynamic support system loading mechanisms: (1) and (2) for loading by
strainburst (SB) and (3) for loading by remote seismic event (modified background figure
from Hoek et al. (1995)).
Furthermore, if the internal energy sink provided by the rock rein-
forcement is ignored, SEbolt plate =aB Esystem. In this case it is as-
sumed that the support system is directly loaded and Esystem = SEbolt
plate + Esurface = (aB + aS)E system. The system capacity is the cumula-
tive capacity of the direct loading capacity of the bolting and the
capacity of the surface support system. In other words, the load
split between the bolting and surface support has to be established
(see Sections 5.2.1 and 5.2.2).
In strainbursting ground, the burst volume VSB fails inside or im-
mediately behind the reinforced rock mass, hence loading the sup-
port in several ways:
A. By accelerating a wedge (or volume of fractured rock) between the
bolts (red in Figure 5-2) and transferring the wedge energy via the
retention system to the bolts (red arrows). Displacements from rock
mass bulking may also directly load the bolt heads (plates; as shown
by red arrows).
B. By accelerating wedges (or volume of fractured rock) between the
bolts (green) and transferring the wedge energy via the retention sys-
tem to the bolts. The bulking strain inside the burst volume will indi-
rectly load the bolts via relative displacements of the fractured rock,
i.e., straining of the bolts inside the burst volume.
In this case, the damaging energy is released from the rock mass
surrounding the burst volume.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 177

When loaded by a stress wave from a remote seismic event, the


energy released at the seismic source causes three (simultaneous)
damage processes. The stress wave loads the reinforce rock mass
and causes damage by:
C. (a) impact loading of the support arch (producing a dynamic tangen-
tial stress flow around the excavation and straining of the reinforced
rock mass; blue arrow indicating tangential hoop stresses);
(b) stress wave reflection at the excavation boundary causing tensile
spalling; and
(c) momentum transfer of some of the source energy to blocks or
wedges near the excavation wall.
The tangential loading mechanism (C.a) may trigger a strainburst.
Ground velocity magnifications near the excavation boundary may
enhance the two sub-mechanisms (C.b) and (C.c).
As a consequence, depending on the energy source and the result-
ing failure process, different percentages of dynamic loading will

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
be shared among the rock mass, the bolts, and the retention system.
Intuitively, it would seem that much energy would be consumed
by breaking rock and by deforming the reinforced rock mass.
However, for support design purposes it is frequently and typically
assumed that no energy is dissipated by the rock mass and the rock
reinforcement. Both Erock mass and SEreinforcement are ignored, i.e.,
Esystem = SEbolt plate + Esurface. This is a conservative assumption as
long as the surface support does not fail prematurely (weakest link
issue).
In the following discussion, energy dissipation and support load-
ing is explored separately for each of the three processes ((A) to
(C)) introduced above and the energy usage is partitioned by the
various energy sinks.

5.2.1 Support to mitigate rockburst damage dominated by dy-


namic disturbances from remote seismicity (direct loading)
For the situation of loading by a dynamic impact or energy trans-
fer (Situations (A) and (B)), some useful insights into a system’s
capacity can be obtained from the South African and the Swiss
test setups shown in Figures 4-14 and 4-15. For the South African
test setup, the drop weight impacts an ‘impact-plate’ placed on top
of concrete blocks (resembling broken rock in the system) and
provides the input energy to the rock and rock support system.
This input energy loads the bolts via the retention system and no
energy is directly transferred to the bolts (Scenario 1 in Figure
5-2). By breaking the concrete blocks and deforming the surface
support (Esurface), a portion of the input energy is absorbed. Most
importantly, part of the energy (Ebolt plate) is transferred to and con-

Draft manuscript – Copyright protected – Cai and Kaiser 2018


178 Rock support system capacity

sumed by the yielding bolts loaded via the surface support and the
plates.
For example, during a test of the ‘Brunswick’ support system,
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

consisting of six 16 mm conebolts, three 0#-gauge straps, mesh


and concrete blocks, 42 kJ of input energy was supplied by a drop
weight (3580 kg dropped from 1.2 m (Ortlepp and Erasmus 2002)).
The surface support consisting of standard mild-steel mesh (100
mm aperture, 4.9 mm wires) and three mesh straps (300 mm wide
with 4 strands, 100 mm aperture, 8 mm wires) deformed about
120 mm without visible damage. Because part of the input energy
was absorbed by the concrete blocks and six conebolts, the portion
of energy absorbed by the surface element can only be obtained if
the three energy sinks (broken concrete/rock, bolts, and surface
elements) can be separated.
Assume that when 25% of the energy is consumed by breaking the
concrete blocks and all six 16 mm conebolts yield at 100 kN (200
kN for 22 mm diameter conebolts), the energy consumed by the
surface system Esurface can be calculated and is presented in Figure
5-3. The energy consumed by the surface system drops rapidly,
i.e., to less than 10 kJ at more than 36 mm bolt displacement (< 18
mm for the 22 mm bolts). Alternatively, if bolts with the same ca-
pacities are displaced by only 10 mm (e.g., a rebar), the surface
element would have to consume between 20 and 25 kJ for 16 and
22 mm bolts, respectively.
50

45

40

35 E(surface) 16 mm

30 E(surface) 22 mm
Energy (kJ)

25

20

15

10

0
0 10 20 30 40 50 60
Bolt displacement (mm)

Figure 5-3 Energy of surface component as a function of bolt yield displacement; exam-
ple for 16 and 22 mm conebolts with 42 kJ input energy and 25% energy loss in
rock/concrete crushing.
Clearly, it is inappropriate to assume that the input energy in these
tests is equal to the energy capacity of the surface support being
tested. Esystem is only equal to Esurface when extremely stiff bolts

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 179

with an infinite capacity are used in a test system or the connec-


tion to the bolts represents the weakest link. This situation is tested
in the GRC setup (Figure 4-13) where the load directly impacts
the surface support element suspended by non-yielding, stiff hold-
ing elements.
If deformable bolts are used, the bolts will dissipate a portion of
the applied input energy and the total energy capacity of the sys-
tem Esystem = SEbolt plate + Esurface. Using yielding bolts in a support
system is therefore an effective means to increase the support sys-
tem capacity for two reasons: more deformable bolts consume
more energy and help prevent premature failure of the connections
as they limit the impact forces at the plates. As a result, the mesh
is less likely to be damaged by the sharp edges of the plates and
the bolt threads are less likely to fail.
However, excessive yield will facilitate rock mass loosening and
bulking, and this can cause shear failures of bolts and excessive
straining of rebar or cablebolts if used in parallel with yielding

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
bolts.
The above statements are only valid when rock directly impacts
the retention system and transfers the impact energy to the surface
component. It is not valid when part or all of the impacting rock
directly loads the bolts. In that case, a lesser part of the impact en-
ergy will be dissipated by the surface system (see next section).
For bolt design purposes, a conservative conjecture would be to
assume that all the energy is used up by the bolting system alone
and, vice versa for surface component design, it would be con-
servative to assume that all energy is dissipated by the surface
support component. However, both assumptions may lead to unre-
alistic and uneconomic solutions and it is necessary to anticipate
the interaction between rockbolts and surface retaining compo-
nents to establish the combined capacity of the integrated support
system.
Bucher et al. (2013) used the Swiss system shown in Figure 4-15
and estimated the percentage of energy split between rockbolts,
high strength mesh, and the test frame using their drop test results.
They found that the split of the energy depends on the type and
stiffness of the holding elements as well as the rock mass behav-
iour (simulated by a concrete slab above the mesh). The total en-
ergy input to the system was 200 kJ. In a stiff slab setup, reflecting
distributed mesh loading, the test system absorbed 158 kJ by
breaking the concrete slab and compacting the rock boulders. Of
the total energy input, 16% or 32 kJ was consumed by four D-
bolts (20 mm diameter, 1.5 m stretch length, split-tube suspension)
and only 5% or 10 kJ was absorbed by the mesh. In this case, of
the total energy absorption by the support system, three quarters
(75%) came from the bolts and one quarter (25%) from the mesh.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


180 Rock support system capacity

In the softer setup using a soft concrete slab, representing more


localized (point) loading, the energy distribution was 5% or 10 kJ
to the bolts, and 13 % or 26 kJ to the mesh for a total of 36 kJ. In
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

this case, 6 kJ or 3% more energy was used in breaking the ‘rock’


and slightly less than three quarters of the support system energy
dissipation came from the mesh. Only slightly more than one
quarter was consumed by the bolts.
It follows that the energy partitioning between retention and hold-
ing systems can be highly variable, as it depends on the dynamic
failure process and the resulting displacement pattern imposed on
the surface and holding components.
In reality, however, the rock mass is rarely loaded as in the test
setup and stress flowing around the excavation causes tangential
straining of the reinforced rock mass as indicated by the blue ar-
row in Figure 5-2. This can lead to strainbursts with radial bulking
deformation and loading mechanisms as described on the right
side of this figure (1 and 2; discussed in the next section).

5.2.2 Support to mitigate strainburst damage dominated by


bulking and shear rupture (direct and indirect loading)
Rock support systems fulfil a dual role:
- at the surface of an excavation, they retain broken rock and
transfer the load via the plates to the bolting system; and
- internal to the rock mass, they have to survive imposed extension
and shear strains caused by the sudden bulking of the burst vol-
ume.
The second mechanism is illustrated by Figure 5-4. Loading via
the retention system or direct loading by dynamically failing rock
(1) imposes a force on the bolt head (blue arrow). Point anchored
bolts, conebolts or debonded cables with anchors, cones or bonded
sections behind the burst volume are loaded in this fashion. On the
other hand, fully grouted bolts or bolts with multiple anchors (e.g.,
rebar, D-bolts, etc.) within the burst volume experience mostly
indirect loading (2) by deep-seated straining of these bolts (green
double arrow in Figure 5-4). This internal straining can be uniform
(linear red line) if bulking is evenly distributed (constant bulking
factor) or, more likely, step-like when strain localization occurs
due to the opening of fractures (indicated by red dashed line). Pre-
viously fractured rock between the burst volume and the excava-
tion wall may be compacted (indicated by the declining red dashed
line).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 181

Figure 5-4 Strainburst loading of support with internal straining in strainburst volume and
direct loading at plate (by force F).
In Figure 5-4, dfo stands for the depth of failure before a strain-

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
burst with a depth of dSB occurs. During the strainburst, the strain-
burst volume bulks and deforms toward the excavation dSB. This
bulking deformation occurs over a time increment Dt, called rup-
ture time (increment), and this causes an initial velocity of vi =
dSB/Dt. It is called ‘initial’ velocity because an effective support
system eventually lowers the velocity to zero when a new equilib-
rium is established. The initial velocity represents the initial im-
pact velocity on the rock in front of the burst volume (the ‘bur-
den’). If the rock is stress-fractured and compressible, the impact
velocity on the support plate in front of the ‘burden’ is less than vi.
In strainbursting ground, only a small part of the energy released
from the ground surrounding the burst volume is dissipated by the
failing rock. The burst volume is an imploding seismic source that
dynamically expands and strains the support elements through the
sudden bulking process in a direct and indirect manner. In this
case, it is reasonable to set Erock mass = 0 and to design a strainburst
resistant support system that provides the other three energy sinks:
a) Ereinforcement: bolts providing energy dissipation capacity by resisting
internal indirect loading;
b) Ebolt plate: bolts providing energy dissipation capacity by resisting
direct loading at the plate; and
c) Esurface: surface support providing energy dissipation capacity if, and
only if, the initial velocity causes rock blocks between the bolts to
directly impact the retention system.
For a support design in burst-prone ground, it is necessary to es-
tablish the energy sharing split between these three possible ener-
gy sinks: Esystem = Ebolt plate + Ereinforcement + (0 to Esurface).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


182 Rock support system capacity

The multipliers aB and aS, as demonstrated by Bucher et al. (2013)


and discussed above, depend on the rock mass stiffness and the
type of rockbolts and surface support used. For the above present-
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

ed examples, aB ≈ 0.75 and aS ≈ 0.25 for stiff loading and aB >


0.25 and aS < 0.75 for soft loading, respectively. According to
Bucher et al. (2013), for the system tested, the mesh dissipated
approximately 70% of the impact energy or aS ≈ 0.7. According to
the South African test results, the mesh dissipated about 25%; thus
aS ≈ 0.25. Even if more case histories were available, it would
likely be impossible to give specific values for a B and aS because
the impact process and load sharing is highly variable. The follow-
ing ranges therefore represent practical ranges: 0.75 > a B > 0.25
or vice versa 0.25 < aS < 0.75.
According to Kaiser et al. (1996), for central deflections in the
range of 75 to 125 mm, mesh-reinforced shotcrete can dissipate
about 3 to 5-times more energy than mesh alone. However, mesh-
reinforced shotcrete panels represent a stiffer surface support sys-
tem and it must be assumed that the load will be effectively trans-
ferred to the bolts. Due to a lack of measurements, the load split
factors cannot be obtained for these tests. In the extreme, the en-
tire load could be transferred to the bolts (aB = 1). On the other
hand, if the mesh-reinforced shotcrete dissipates between 15 and
20 kJ (see Chapter 4) before being excessively damaged, the sur-
face support system factor could be as high as aS = 0.5. Accord-
ingly, the following range may be applicable for mesh-reinforced
shotcrete: 0.25 < aS < 0.5 or 0.75 > a B > 0.5. It is reasonable to
assume that bolts will attract more energy when connected to
mesh-reinforced shotcrete. Similarly, it must be assumed that bolts
will attract more energy when heavy straps over mesh are used.
As indicated above, assuming a R = 0 could lead to an excessively
conservative design. It is necessary to better understand the inter-
action between the rock and the rock reinforcement system as
much energy can be dissipated by the reinforced rock mass rather
than by the support alone. Further research is required to advance
the knowledge concerning the dynamic self-supporting capacity of
a supported rock mass. As Hoek and Brown (1980) mentioned for
static rock support design, support is needed to enhance the self-
supporting capacity of the ground. The same applies to burst-
prone ground. It is the role of the support to enhance the dynamic
self-supporting capacity of the rock mass which entails three as-
pects: (1) to raise the rock mass strength near the excavation to
prevent burst damage (if possible); (2) to minimize the energy re-
lease from the failing/strainbursting rock near the excavation; and
(3) to consume all released energy in the reinforced and supported
rock mass (see discussion on the ‘gabion concept’ below).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 183

5.2.3 Rock support system testing – simulated rockbursts


The previously described test facilities were designed to assess the
capacity of individual or interacting support elements. Because a
rock support system is comprised of various support components
such as different rockbolts, mesh, and shotcrete that interact dif-
ferently with the ground, it is important to establish the capacity of
various support systems.
Unfortunately, no existing test facility can dynamically test the
capacity of entire rock support systems consisting of various com-
ponents with rock-support interaction. Although the WASM facili-
ty is able to test parts of a ‘rock support system’ (Villaescusa et al.
2016), the rock support system is limited in size and number of
bolts that can be installed. The South African and the Swiss drop
test facilities (Stacey and Ortlepp 2001; Bucher et al. 2013) are
capable of testing part of a rock support system comprised of
rockbolts, surface support elements (mesh-, or fiber-reinforced
shotcrete), and broken rock (represented by concrete bricks).

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
However, most published test results focus only on the evaluation
of surface support capacity.
During field testing, it is difficult to properly represent the dynam-
ic loads and deformation fields experienced by the reinforced rock
masses during a rockburst. Furthermore, it is difficult to define
and control the nature of each possible rockburst damage mecha-
nism. These difficulties can partially be overcome by simulated
rockbursts using blasting. However, because a blast-simulated
rockburst is not real, it has to be classified as a pseudo-rockburst
experiment.
Blast-simulated rockbursts do not represent the prevailing condi-
tions imposed by the energy release from major remote seismic
events. Such events modify the stress field surrounding the entire
excavation, cause cyclic stress changes, and the damage is rarely
related to high frequency ground motions created by blasts. Blast-
simulated rockburst tests are most representative of strainbursts
but are plagued by one major drawback, i.e., the gas pressure not
experienced during strainbursts.
Despite the fact that such tests are flawed in many ways, they do
provide insight into the dynamic rock–support interaction process.

Simulated pseudo-rockburst experiments


The first pseudo-rockburst experiment to study the in-situ perfor-
mance of a rock support system was conducted in South Africa in
1969 to demonstrate the benefit of yielding bolts (Ortlepp 1969,
1992). The test was carried out at an intersection that allowed the
drilling of blast holes parallel to the test walls. 1.8 m long conven-
tional bolts (rebar) and yielding bolts were installed on the left-

Draft manuscript – Copyright protected – Cai and Kaiser 2018


184 Rock support system capacity

and right-hand sides of a tunnel, respectively, along with double-


layer 8-gauge linked 50 mm wire mesh (Figure 5-5). The yielding
bolts were smooth bars with a 22.5 cm long threaded portion at the
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

anchor end. A smooth-bored die of internal diameter less than the


threads was fitted to the bolts. Slippage occurred at a constant load.
Blast holes, 3 m long at 43 cm spacing, 60 cm from the wall
boundary, were charged to generate dynamic ‘loads’. The rock
and bolt responses were not monitored to quantify the capacities
of the two support systems.

Conventional bolts

Yielding bolts

Blast holes

Tunnel boundary
after blast

Original tunnel boundary

1m

Figure 5-5 Field dynamic experiment to compare the support capacities of yielding bolts
to conventional bolts (reproduced after Ortlepp (1969)).

The dynamic ‘load’ in this experiment basically consisted of two


overlapping damage processes: (1) a stress wave radiated and
caused tensile stresses at the wall that may have produced exten-
sion fractures; and (2) the volume of the blast-damaged rock in-
creased rapidly imposing a displacement pulse on the ‘burden’, i.e.,
the rock in front of the blast volume.
After the blast, the left side of the tunnel supported by the conven-
tional bolts failed whereas the right side supported by the yielding
bolts survived the blast’s dynamic load (Figure 5-5). This experi-
ment, with surface parallel blast holes in close proximity to the
excavation surface, simulated rock ejection by an explosive im-
pact velocity at the blast ring location. It simulated the sudden
bulking impact released by a strainburst1 and demonstrated in a

1
At the time, researchers thought that this type of experiment would represent rockburst
conditions caused by large remote seismic events. Today, it is understood that this type of
experiment is more closely related to conditions created by strainbursts (with the excep-
tion of gas pressure effects).
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 185

rather convincing manner the superior capacity of yielding-


rockbolt-based support systems to dissipate energy transferred to a
mass of rock near the excavation wall.
Other simulated pseudo-rockburst experiments have been con-
ducted in underground mines in Canada at Campbell Mine
(Hedley 1992), Bousquet Mine (Tannant et al. 1995; Kaiser et al.
1996), Inco’s 175 Orebody (Espley 1999), Fraser Mine (Andrieux
et al. 2005), in South Africa at Kopanang Mine (Hagan et al. 2001;
Haile and Le Bron 2001), in Chile at El Teniente Mine, in Austral-
ia at Long Shaft Mine (Heal et al. 2004; Potvin and Heal 2010),
and at Kiruna Mine in Sweden (Shirzadegan et al. 2016).
The blast holes at some sites (e.g., Kopanang Mine and Long
Shaft Mine) were located in deep ground well behind the anchor-
ing points of the rockbolts, such that rock damage near the tunnel
surface was primarily caused by dynamic shock waves.
At other sites, the blast holes were located between the anchoring
points and the surface plates (e.g., as shown in Figure 5-5). In

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
these conditions, the bolts were indirectly loaded by straining near
the blast hole ring and directly loaded by impact forces at the plate.
Inclined blast holes were used at the Bousquet Mine and Fraser
Mine test sites to simulate various loading types at various depths
of strainbursting. As indicated above, blast tests inside the sup-
ported volume are most representative of strainbursts and the as-
sociated mixed loading mechanisms consisted of dynamic stress
waves emitted from a local source and a simulation of rock mass
bulking (even though the latter may be exaggerated by the influ-
ence of gas pressures). With the inclined holes, the entire spectrum
between stable and failing support was investigated in one exper-
iment. For detailed interpretations of such an experiment, the
reader is referred to Section 4.10.3 of Kaiser et al. (1996).
There are fundamental differences between blast and rockburst
damage processes. The seismic waves generated by a remote
seismic event cannot be compared with those generated by a blast
as the resulting stress and displacement fields differ in many ways.
The typical signature of a seismic event is initiated by a relatively
low frequency compressive p-wave followed by a larger amplitude
shear wave. On the other hand, blasts produce mostly p-wave
ground motions, particularly if the detonation velocity is very high
(Hildyard and Milev 2001), and shockwaves with higher frequen-
cies (Hadjigeorgiou and Potvin 2007). It follows that blast exper-
iments with near-wall blast locations cannot be used to simulate
damage processes caused by large remote seismic events. Such
simulated rockbursts also cannot reproduce continuous or repeated
shaking effects leading to unravelling or shakedown failures.
However, blasts cause local stress fracturing and driven by gas
pressure simulate local rock mass bulking (even though this may

Draft manuscript – Copyright protected – Cai and Kaiser 2018


186 Rock support system capacity

be exaggerated if the gas pressure does not dissipate through


stress-fractured ground). As a consequence, the test results ob-
tained from simulated pseudo-rockbursts are indicative of failure
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

processes encountered during self-initiated, triggered or dynami-


cally loaded strainbursts. The results define in a qualitative man-
ner the response and capacities of various support systems to
strainbursting. Interestingly, the experiments at Bousquet Mine
(Tannant et al. 1995; Kaiser et al. 1996) demonstrated an often
observed characteristic of strainburst damage, i.e., a rapid transi-
tion from stable to collapsed ground support.
A summary of various dynamic testing methods is presented in
Figure 5-6. Each method is useful to find specific characteristics
of rock support elements and behaviours of rock support systems.
Because of differences in test boundary conditions, care must be
exercised when comparing test results obtained from various drop
test facilities and from field tests using blasting techniques. Drop
tests and simulated rockbursts are best used to compare the rela-
tive performance of support elements or systems under one given
test condition. Ground truthing then serves to validate the capacity
of a chosen rock support system against various demands imposed
by actual rockbursts.

Dynamic Tested
Test method Advantage Disadvantage
loading capacity

Direct impact / Well defined Rock medium is


Support loading, missing, limited to
Lab drop test Moment
element repeatibility, low dynamic impact
transfer cost loading

Can control testing


Loading only
Simulated location and timing,
representative of
Blasting Support system test of rock support
rockburst system in actual
strainbursts, time
consuming, costly
ground

Dynamic Real ground


Uncontrolled timing,
Ground condition, real
bulking or Support system test location, and
truthing loading condition,
event size
seismic wave real system testing

Figure 5-6 Summary of dynamic test methods.

5.3 Estimation of support system capacity


Much effort has been expended around the world to establish di-
rect and indirect load, displacement, and energy dissipation capac-

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 187

ities of support components (Chapter 4 and Appendix E). Howev-


er, there is no systematic engineering approach to estimate the ca-
pacity of support systems including the bolting system and the sur-
face support. In the following, we offer a proposed means to esti-
mate the direct load capacity of support systems. In this approach,
the dynamic self-supporting capacity of the reinforced rock mass,
including the capacity of the reinforcement and the rock, is ig-
nored (i.e., aR = 0). The support system capacity is Esystem = SEbolt
plate + Esurface. with SEbolt plate representing the cumulative energy
capacity of all installed bolts at a given mining-induced displace-
ment dbolt plate. A major difference of this approach from the exist-
ing approaches is that energy sharing between support components
is considered and the deformation compatibility is respected for
the entire rock support system. In this manner, the false approach
of summing the maximum tested component capacities to obtain
the system capacity is eliminated.
This approach implies that there are no weak links. All connec-

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
tions are sufficient to survive the dynamic impact loads, and the
tendon and surface energy capacities are simultaneously mobilized
and dissipate energy in parallel. Furthermore, this approach re-
spects that the support capacity can be consumed as the support
system is being deformed.

5.3.1 Support capacity consumption


Bolting system
Mining causes not only stress changes but also associated defor-
mations and tunnel convergence (wall displacements) which de-
form and strain the support. As these displacements increase, part
of a support’s displacement or energy dissipation capacity gets
consumed. This is called support capacity consumption and is re-
flected by the support energy consumption plot shown in Figure
5-7. The support system capacity for this illustrative example is
gradually lost until 100% of its capacity is consumed at 200 mm
imposed displacement (dashed curve). If this support system were
enhanced between 100 and 120 mm wall displacement by adding
long plain cables, about half of the consumed capacity would be
restored (red line).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


188 Rock support system capacity

100%
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Energy capacity consump9on (%)


80%

60%

40%

Enhanced with cables at 120 mm


20%
Original support

0%
0 50 100 150 200 250
Displacement imposed on bolt system (mm)

Figure 5-7 Illustration of energy capacity consumption by support deformation and sup-
port system capacity restoration by proactive support maintenance.
As capacity is consumed, a support system has less remnant ca-
pacity to resist displacements imposed by future dynamic disturb-
ances. For the example shown in Figure 5-7, 70% of the energy
capacity of the original support is consumed at 100 mm wall dis-
placement. However, if the support is enhanced with cables, 35%
of its original energy capacity is restored, thereby increasing its
potential ability to survive a future dynamic event (e.g., strainburst)
even if this additional capacity is eventually also consumed at 200
mm wall displacement for this case.
Retention system
Similarly, the energy capacity of retention systems can be con-
sumed as they are deformed. For example, mesh-reinforced shot-
crete (according to Figure 4-20b) would have consumed between
10 and 15 kJ/m2, and 15 and 20 kJ/m2 for central deflections of
100 and 200 mm, respectively. The mesh-reinforced shotcrete
would, at this stage, be moderately damaged, leaving, at best, an-
other 15 kJ/m2 remnant energy capacity before failure, which
would be expected at about twice the above quoted displacement
range.
In other words, it must be assumed that E100 to E200 represents a
meaningful range for design capacities. The remnant capacities are
comparable but get gradually eroded to zero at the point of col-
lapse. For design purposes, it is to be assumed that half the energy
capacity is consumed at a threshold displacement dS (= 100 to 200
mm for mesh-reinforced shotcrete) and the other half when 2dS is
reached.
If the energy capacity of a support system can be consumed, it fol-
lows that part of it can be restored by installing additional support
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 189

after some displacements have been imposed on the system. This


is called proactive or preventive support maintenance (PSM).

5.3.2 Proactive support maintenance


In burst-prone ground, it is necessary to maintain at all times a
sufficient energy or displacement capacity margin of the support
system to survive the impact of rockbursts. This may be achieved
by providing sufficient ductility (using yielding bolts). However, it
is often not possible to provide sufficient ductility to the entire
support system, e.g., due to the unavailability of sufficient num-
bers of yielding bolts, the depth of failure exceeding the length of
the yielding bolts, an inability to debond long holding elements
(e.g., installed cables), or because the impact of mining has con-
sumed more support capacity than originally estimated. In these
situations, the example presented above demonstrates that the
burst resistance of a support system can be increased by proactive
support maintenance (PSM).

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
The concept of support system capacity consumption is further
illustrated by Figure 5-8 with photos demonstrating increasing
support consumption toward the location where the support and
the excavation eventually failed (at the back end of the drift). This
figure also highlights the displacement range where the original or
baseline design is valid (e.g., to the displacement limit of about
120 mm at [2] in the figure) and when proactive support mainte-
nance is needed or most effective (at between 120 and 200 mm,
between [2] and [3] in the figure). The displacement scale depends
on the composition of the integrated support system. If the oppor-
tunity is missed to proactively enhance the support, failure may
occur and rehabilitation2 will be required (beyond [3]) as will be
discussed later.
Proactive support maintenance is a practical and often an econom-
ical means to increase workplace safety and reduce the potential
severity of rockburst damage. This is particularly meaningful
when deep-seated failure is expected and yielding bolts are too
short to reach stable ground or debonded cables cannot be in-
stalled. PSM is particularly beneficial when unexpectedly large
convergences are encountered. It also provides a viable alternative
to yielding support when the impact of an advancing stress front
(e.g., undercut advance in block caving) with deep seated rock
mass deformations has to be managed. PSM may be more eco-
nomic than installing burst-resistant yielding support systems

2
It is important to distinguish between support ‘maintenance’ and ‘rehabilitation’. Sup-
port maintenance means that the support is upgraded to ‘maintain’ sufficient capacity
during future rockbursts. Support rehabilitation implies that the support was damaged to
the point where it needs to be replaced to ‘restore’ the desired support capacity.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


190 Rock support system capacity

across the board, particularly when excessive support demand is


localized and the locations requiring such support are not a priori
foreseeable. The examples presented in Section 5.2.2 demonstrate
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

how PSM can extend the displacement capacity of a support sys-


tem and restore a substantial part of the systems’ energy dissipa-
tion capacity.
As is discussed in Volume II, deformation-based support selection
and PSM procedures can be developed utilizing mine-specific
convergence measurement data.

Base PSM
Design Rehabilita/on

Figure 5-8 Illustration of support system capacity consumption and range of applicability
of base design, proactive support maintenance (PSM) and support rehabilitation (Photo
courtesy: Deep Mill Mining Zone at Grasberg Mine, PT Freeport Indonesia 2017).

Maximum practical displacement limit (MPDL)


Rockburst damage often leads to extensive rock fracturing and as-
sociated rock mass bulking. For design purposes, it is therefore
necessary to define the maximum ‘allowable’ inward displace-
ment (wall convergence) that can be permitted before drifts be-
come functionally disabled or good quality support installations
become compromised. The maximum allowable inward displace-
ment depends on the robustness of the installed rock support sys-
tem and the space tolerance (initial drift width minus equipment
size).
Most well-designed and high quality deformable rock support sys-
tems work well and do not lose their overall integrity within a one-
sided radial displacement range of 200 to 300 mm. This includes

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 191

the displacement at the bolts dB and the central deflection of the


surface support dS between the bolts. For support design purposes,
dB =150 to 200 mm and dS = 100 to 150 mm for a total of 250 to
350 mm represents a reasonable maximum practical displacement
limit (MPDL).
The photographic examples presented in Figure 5-8 show one-
sided drift closures that exceed 300 or even 400 mm after a large
seismic event. It can also be seen that the support’s functionality
has been severely compromised where larger displacements were
imposed on the support system. Whereas temporary access may
have been preserved, rehabilitation is required in situations where
the MPDL is exceeded.
The MPDL guides the selection of practically meaningful capaci-
ties for support components as obtained from laboratory tests or
from manufacture specifications. Selected design parameters must
be consistent with the chosen allowable or practical displacement
range or the MPDL for the bolting and surface support. This is il-

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
lustrated in Figure 5-9 by hiding capacities that cannot be obtained
before the practical displacement limit is reached.
By restricting the applicable displacement range to 150 mm (red
rectangle; full line) each for the bolting system and the central de-
flection between the bolts for the surface support, the following
maximum energy capacity limits (MEL) apply:
- Individual bolts: MEL = 5 to 32 kJ/bolt; and
- Surface elements: MEL ≤ 12 kJ, i.e., for impact areas ranging
from 1 to 1.5 m2, MEL = 5 to 12 kJ/m2.
If both contribute equally to the wall displacement, the combined
MEL = 10 to 44 kJ/m2.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


192 Rock support system capacity
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

(a) Bolting elements

(b) Surface support elements


Figure 5-9 Limited range of practically acceptable support capacities; data outside the
practical range are shown in a transparent fashion (data compilation by Potvin et al.
(2010) from data by Kaiser et al. (1996), Stacey and Ortlepp (2001), Gaudreau et al.
(2004), Falmagne et al. (2005) and Player et al. (2008)).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 193

The data in the transparent section suggest that individual compo-


nents may have extra capacity at displacements exceeding the
MPDL. Whereas this is the case in laboratory testing situations for
individual components, the ultimate capacity does not matter as it
cannot be reached in an effective support system. On the contrary,
whereas low energy values at a displacement below the MPDL
suggest that the support fails and does not dissipate energy, this is
misleading as the respective energy components actually contrib-
ute to the system’s energy dissipation capacity as long as the sup-
port system integrity is maintained (with dB + dS < MSPL).
After proactive support maintenance, however, some of originally
installed support components may still have extra energy dissipa-
tion capacities as highlighted by the red dashed rectangular area in
Figure 5-9 (for a further 150 mm of bolt and surface support dis-
placement). They can be mobilized if the integrity of the support
system is maintained or re-established by PSM. In such situations,
it may be possible to maintain the integrity of a support system to

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
a total cumulative wall displacement of 0.6 m.

5.3.3 Direct impact loading capacity of bolting system


Based on the above discussion, it is conservatively assumed that
the capacity of the bolting system is equal to Ebolt plate. In other
words, the reinforcement and the rock mass energy dissipation ef-
fects are ignored when such an assumption is made. This is con-
servative because there are always some internal or indirect load-
ing processes that consume energy inside the burst volume. The
bolting system capacity can then be combined with the surface
system capacity, as discussed in Section 5.2.2, as long as there are
no weak links between the bolting and surface support. Means to
estimate the indirect energy consumption capacity of rock rein-
forcements will have to be developed in the future (a preliminary
solution is presented in Volume II).
Considering the displacement characteristics shown in Figure 5-10,
it is reasonable to approximate the load–displacement relation of a
bolt by an equivalent perfectly plastic bolt model (red lines). This
energy capacity is defined by the product of the ‘average’ or mean
load capacity Fm (see Table 5-1) and the allowable displacement
capacity dall.
For the test curves presented in Figure 5-10, these capacities are
presented in Table 5-1. The single cable capacity is based on Fig-
ure 5-11.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


194 Rock support system capacity
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Figure 5-10 Approximations of load–displacement relation by perfectly plastic model


(red capacity lines) for four examples (modified Figure 15 from Li (2010)).

Table 5-1 Allowable design capacities for bolting system components chosen for demon-
stration purposes
Bolt type(I) Mean force Allowed(III) displace- Allowed energy capacity
Fm (kN) ment capacity dall (mm) Eall = Fm*dall (kJ)
Resin rebar 160 30(II) 4.8

Cement rebar 170 40(II) 6.8


(3–8)(IV)
Spilt set 50 150 7.5
(2–8)(IV)
Super Swellex 90 150 13.5
(10–35)(IV)
D-bolt (1 m paddle length; 175 150 26.3
indirect load) (26–38/m)(IV)
Direct loading at plate 0.4 m 67 11.7
to first paddle anchor (10–15) (IV)
Single plain cable 225 50 11.3
at W:C ≤ 0.35 (5–9) (IV)
(I)
Bolt dimensions are omitted because the numbers in this table are only used as
illustrative examples (for recommended values refer to support capacity tables in
Chapter 4).
(II)
Values may be less, as low as half, for direct loading conditions if the bolts
are perfectly grouted to the plate and the rock is not relaxed near the wall.
(III)
The allowed displacement capacity and the corresponding energy capacity

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 195
may differ from those obtained from field pull-out tests3.
(IV)
The ranges listed in Tables 4-2 and 4-3 for E100 or E200 are shown in
brackets ().
These values are not generally applicable design parameters as
they may vary widely depending on the supplier and on-site condi-
tions. They are used in the following to calculate the bolting sys-
tem capacities of illustrative examples.

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
(a)

(b)
Figure 5-11 Approximations (red lines) used for direct loading of cable; load–
displacement charts from the Cable Bolt Handbook (Figure 2.91 for plain cables in
Hutchinson and Diederichs (1996)).
For the estimation of the load, displacement, and energy capacities
of a bolting system, it is necessary to make an assumption on how
and when individual components are displaced. It is assumed for
the following discussion that all bolts work in parallel, meaning

3
Pull-out tests may have been terminated prematurely due to equipment limitations and
safety concerns and dult may actually be larger than what is considered to be practically
allowable for a design. The practical dult may therefore be less than that obtained in tests
depending on the holding process; e.g., a Split set may be loaded for half of its length and
the allowable displacement is much less than dult obtained from pull-out tests on full
length bolts.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


196 Rock support system capacity

that they are simultaneously deformed once installed (in direct or


indirect loading). In reality, these bolts are differentially deformed
as not all bolts get installed at the same time, and the installation
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

sequence varies. This has to be considered when assessing the


bolting system capacity. In other words, identical bolts installed at
different stages of mining will not reach their peak or ultimate ca-
pacities at the same ‘time’.
For the illustrative examples, the bolt installation sequence pre-
sented in Table 5-2 is specified. The third column shows the
amount of wall or plate deformation preceding the installation of a
particular bolt at the time of bolt activation and the fourth column
provides the assumed bolt densities (square spacing).
This table summarizes a situation where Split sets are used during
the development to pin the mesh and a secondary bolting pattern
of intermittent rebar and D-bolts is installed after 10 mm conver-
gence has occurred. Single plane cables are added after 40 mm
wall displacement is observed and proactive support maintenance
is conducted with single plane cables after 75 and 125 mm of wall
displacement are recorded.
Table 5-2 Allowable design capacities for support system capacity assessment
Bolt type Bolt # (I) Plate displacement before Average # bolts/m2
as per Table 5-1 bolt activation (mm) (spacing in m)

Split sets 1 0 0.75 (1.2)


Resin rebar 2 10 0.5 (1.4)
D-bolt 3 10 0.5 (1.4)
4 40 0.5 (1.4)
Single plated
5 Rehab at 75 mm 1.0 (1.0)
plane cables
6 or at 125 mm 1.0 (1.0)
(I)
Bolt # refers to numbering in the support capacity estimation spreadsheet.

Means to estimate the surface support capacity for integration into


the support system capacity will be discussed in Section 5.3.3.
These surface support capacities are superimposed in the follow-
ing figures for illustrative purposes.

Example of direct loading capacity of bolting system


An Excel spreadsheet was developed for the purpose of exploring
alternate bolt combinations. It can be downloaded at the following
link ‘Support system characteristics-v1803.xlsx’ (user instructions
are provided in Sheet 1, input and output table in Sheet 2, and
graphs in Sheets 3 and 4). This prototype version makes several
simplified assumptions; a more sophisticated version may be re-
leased in the future.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 197

For the following example, it is assumed that the base design con-
sists of the reinforcement package described above, consisting of
Split sets with alternating rebar and D-bolts and plain cables to
hold the reinforcement package in place. The accumulated dis-
placements at the time of bolt installation are 0 mm for the Split
set (at the development face), 10 mm for the rebar and D-bolts (a
few rounds delayed behind the development face), and 40 mm for
the cables (installed after some mining-induced deformations have
been encountered). The respective bolt spacing assumptions for
this base design with the numbers of bolts per square metre are
listed in Table 5-2 (s = 1.2 m for Split sets, 1.4 m for rebar, D-
bolts and cables). The surface support for this example consists of
standard mesh-reinforced shotcrete and it is specified that the cen-
tral deflection between bolts should not exceed 100 mm 4.
The output summary:
Split Set,Rebar,D-bolt at plate,Single Cable 1,
Max. bolting energy capacity 19.0 kJ/m2

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Mesh-reinforced reinforce SC
Max. surface support energy capacity at 100 mm 6.7 kJ/m2
Maximum integrated support capacity 23.8 kJ/m2
Remnant bolting energy capacity at 100 mm 1.6 kJ/m2
Remanant Support system capacity 6.3 kJ/m2
MAX yield force on retention system 225 kN
Max. load capacity 239 kN/m2
Mean load capacity 127 kN/m2
Allowable displacement at 75 kN 90 mm

provides the maximum possible energy capacity of the bolting,


retention and integrated support system assuming that there are no
weak links between the various support components. The remnant
capacity is listed for a specified displacement threshold of 100 mm.
The maximum yield force on the retention system is provided to
facilitate the selection of a retention system with sufficient punc-
ture capacity. The maximum and mean load capacities over the
entire displacement range are provided to assist in assessing the
support system’s capacity to hold the gravitational (static) load
from rock wedges or broken rock. Finally, the allowable dis-
placement is provided for a chosen minimum load capacity (75 kN
in this example).
The corresponding load and energy versus displacement curves for
the bolting and integrated support system are presented in Figure
5-12.

4
The central displacement threshold can be defined by the user.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


198 Rock support system capacity
400
Split Set,Rebar,D-bolt at plate,Single Cable 1,
350 Split Set,
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Rebar,
300
D-bolt at plate,

Load capacity (kN)


250 Single Cable 1,

Single Cable 2,
200
Single Cable 3

150
Cumulative load
Capacity
Average load
100
capacity

50

0
0 50 100 150 200 250
Displacement on bolt system (mm)
(a)
Split Set,Rebar,D-bolt at plate,Single Cable 1,
50
Mesh-reinforced reinforce SC Cummulative bolting capacity
45
Support system capacity
40 Remnant bolting system
capacity
35 Remnant support system
capacity
Energy capacity (kJ/m2)

30

25

20

15

10

0
0 50 100 150 200 250
Displacement on bolt system (mm)
(b)
Figure 5-12 Load (a) and energy (b) capacities of the baseline bolting system with Split
sets, rebar, D-bolts and single cables (with mesh and straps for a central deflection
threshold of 100 mm). The remnant energy capacity after 100 mm is shown in red.
This support system has a maximum and an average load capacity
of 239 and 127 kN/m2, respectively, and it drops below 75 kN at a
displacement of 90 mm. The maximum point load on the retention
system is 225 kN (at the cable plate). The retention system of this
support system would therefore require a very high puncture ca-
pacity that mesh-reinforced shotcrete alone cannot provide unless
very large cable plates or straps are supplied to distribute the load.
Whereas the D-bolts maintain their internal load capacity over a
displacement range of > 150 mm, at the plate, with an assumed
spacing of 0.5 m to the first paddle, this displacement capacity is
reduced to 67 mm. Hence, the holding capacity at the plates drops
after 67 mm displacement and then further when the cables fail at

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 199

90 mm (composed of 40 mm before installation and 50 mm as-


sumed cable displacement capacity). The load capacity drops to
that of the Split sets (37.5 kN/m2 ) when the cables fail at 90 mm.
For a prescribed allowable minimum load capacity5 of 75 kN/m2,
this support system reaches the load capacity threshold at 90 mm
bolt displacement.
The energy capacity of the bolting system measures 19.0 kJ/m2
(Figure 5-12b) but the remnant capacity at 100 mm 6 of wall/bolt
plate displacement amounts to only 1.6 kJ/m2. The Split sets pro-
vide this remnant energy capacity because the cables have failed at
this stage (with < 100 mm wall displacement).
The integrated support system capacity with mesh-reinforced
shotcrete is discussed later. It is shown here (or in Figure 5-12b)
as dashed and dotted lines. The maximum energy dissipation ca-
pacity is 23.8 kJ/m2 at 90 mm of wall displacement before it de-
clines until the support system collapses at 150 mm. A remnant
support capacity of approximately 5 kJ/m2 after 100 mm of wall

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
displacement is largely provided by the surface support.
It follows that proactive support maintenance (PSM) should be
executed for this support system before the plate displacement
reaches the 90-mm threshold, e.g., after 75 mm bolt displacement.

Direct loading capacity of bolting systems after PSM at 75 mm


With a single cable per m2 or a double cable per 2 m2 added after a
plate displacement of 75 mm, the revised load and energy charac-
teristics are presented in Figure 5-13. The load capacity is drasti-
cally increased to an average of 202 kN/m2 and the surface support
will see the same point load of 225 kN. Again, large cable plates
must be installed to provide sufficient puncture capacity.
The total energy dissipation capacity of the bolting system is al-
most 60% higher at 30.3 kJ/m2 and the remnant capacity has more
than quadrupled to 7.2 kJ/m2. This system would provide a thresh-
old load capacity of 75 kN until the cumulative bolt displacement
reaches 125 mm.

5
The allowable minimum load capacity can be defined by the user to ensure that the
bolting support is able to maintain a minimal holding capacity.
6
The displacement threshold to obtain the remnant capacities can also be defined by the
user.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


200 Rock support system capacity

Split Set, Resin bar, D-bolt at plate, Single Cable 1, Single Cable 2,
Max. bolting energy capacity 30.3 kJ/m2
Mesh-reinforced reinforce SC
Max. surface support energy capacity at 100 mm 6.7 kJ/m2
34.3 kJ/m2
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Maximum integrated support capacity


Remnant bolting energy capacity at 100 mm 7.2 kJ/m2
Remanant Support system capacity 11.2 kJ/m2
MAX yield force on retention system 225 kN
Max. load capacity 375 kN/m2
Mean load capacity 202 kN/m2
Allowable displacement at 75kN 125 mm
400 50
Split Set, Mesh-reinforced reinforce SC Cummulative bolting capacity
Resin bar, 45
350 Support system capacity
D-bolt at plate,
40 Remnant bolting system
300 Single Cable 1, capacity
35 Remnant support system
Load capacity (kN)

Single Cable 2, capacity

Energy capacity (kJ/m2)


250
Single Cable 3 30

200 Cumulative load 25


Capacity
Average load capacity
20
150
15
100
10
50
5

0 0
0 50 100 150 200 250 0 50 100 150 200 250
Displacement on bolt system (mm) Displacement on bolt system (mm)

(a) (b)
Figure 5-13 Load and energy capacities of the baseline bolting system plus PSM with one
single cable per m2: (a) load capacity and (b) equivalent energy capacity. The remnant
energy capacity after 100 mm is shown in red.

Direct loading capacity of bolting systems after PSM at 125 mm


If the opportunity to execute PSM at 75 mm was missed and the same
cable pattern was installed at 125 mm, the load and energy characteristics
presented in Figure 5-14 would be obtained. This system would provide
the same total energy dissipation capacity and a further (60%) increase in
the remnant energy capacity of the bolting system to 12.8 kJ/m2.
This support would fail after 175 mm of cumulative bolt dis-
placement. However, due to the late intervention, the bolting sys-
tem experiences a very low load capacity (34.7 kN/m2 ) between
90 and 125 mm of plate displacement. This support is therefore
unsafe during PSM work for two reasons:
- the static factor of safety in terms of load capacity is very low,
rendering the excavation susceptible to shakedown; and
- the remnant energy capacity is very low, resulting in a high vul-
nerability to rockburst damage.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 201
Split Set, Resin bar, D-bolt at plate, Single Cable 1, Single Cable 3
Max. bolting energy capacity 30.3 kJ/m2
Mesh-reinforced reinforce SC
Max. surface support energy capacity at 100 mm 6.7 kJ/m2
Maximum integrated support capacity 31.9 kJ/m2
Remnant bolting energy capacity at 100 mm 12.8 kJ/m2
Remanant Support system capacity 14.4 kJ/m2
MAX yield force on retention system 225 kN
Max. load capacity 260 kN/m2
Mean load capacity 173 kN/m2
Allowable displacement at 75 kN 175 mm

Split Set, Resin bar, D-bolt at plate, Single Cable 1, Single Cable 3
Split Set, Resin bar, D-bolt at plate, Single Cable 1, Single Cable 3 50
400 Mesh-reinforced reinforce SC
Split Set,
45 Cummulative bolting capacity
Resin bar,
350
40 Support system capacity
D-bolt at plate,
300 Single Cable 1, Remnant bolting system
35 capacity

Energy capacity (kJ/m2)


Load capacity (kN)

Single Cable 2, Remnant support system


250 30 capacity
Single Cable 3

200 Cumulative load 25


Capacity
Average load capacity 20
150
15

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
100
10
50 5

0 0
0 50 100 150 200 250 0 50 100 150 200 250
Displacement on bolt system (mm) Displacement on bolt system (mm)

(a) (b)
Figure 5-14 Equivalent system capacities of the support system after proactive support
maintenance with cables added at 125 mm: (a) load capacity and (b) equivalent energy
capacity. The remnant energy capacity after 100 mm is shown in red.

These examples illustrate how the bolting system capacity can be


estimated and demonstrate several important aspects of support
system design for burst-prone ground:
- The load, displacement, and energy capacities are sensitive to
mining-induced deformations.
o They can be greatly enhanced by timely proactive sup-
port maintenance (staged support installation).
- The surface support must at all times provide sufficient puncture
load resistance.
- The remnant energy capacity at a given wall displacement must
be assessed by considering the active support elements.
o In this example, it is almost zero after 90 mm of wall
displacement when the first set of cables fails.
- The timing of PSM is critically important. There is a trade-off to
be considered when selecting the time of PSM. Delayed PSM
adds a displacement capacity and increases the remnant energy
capacity but too much delay may compromise the load and ener-
Draft manuscript – Copyright protected – Cai and Kaiser 2018
202 Rock support system capacity

gy capacities before the PSM is completed. This can lead to un-


safe workplace conditions during the PSM.
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

5.3.4 Contribution of surface support to integrated support sys-


tem capacity
Surface support or retention systems play a critical role in main-
taining the stability of an excavation and in facilitating the gabion
support concept (introduced later in Section 5.3.7). They contrib-
ute in various forms by ensuring good support component perfor-
mance enhancing the support system’s overall capacity and integ-
rity. The surface support:
- prevents unravelling of stress-fractured rock and blocky ground
between bolts;
- transfers the load from broken rock to the bolting system if the
puncture capacity is sufficiently high;
- dampens the direct impact by distributing loads to multiple bolt-
ing elements; and
- helps to minimize rock mass bulking.
For these and other reasons, the retention system must be robust
and all connections must be strong (high puncture capacity) such
that high individual bolt forces cannot compromise the integrity of
the integrated support system.
Whether the surface support actually dissipates part of the energy
demand caused by a dynamic disturbance or only transfers the
load demand to the bolting system depends on various factors as
discussed in Sections 5.2.1. Basically, the surface support only
contributes directly to the energy dissipation if energy is con-
sumed by deforming the surface support. If it channels the entire
demand to the bolting components, the surface support does not
dissipate energy. However, it is rare that all or even the majority
of the dynamic energy is channeled via the surface support to the
rockbolts. As a matter of fact, in practice it is often observed that
the energy capacity of the retention system is exceeded and fails
prematurely, facilitating unravelling of rock between bolts and
ejection of poorly retained rock blocks or fragments.
Retention capacity
A retention system is loaded in two ways, i.e., by broken rock that
is bulking or impacting the surface support between the bolts, and
in the opposite direction by the resisting bolt forces (through the
plates). Both actions – the resistance to impact forces between
bolts and the reaction forces at the plates – can be tested by la-
boratory impact tests. The energy dissipation capacity of the reten-
tion system only matters when it is loaded by broken rock.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 203

Because the surface support works in series with the bolting sys-
tem, it can only dissipate energy from broken rock that is bulking
and dynamically load it between the bolts. Of course, it also helps
the bolting system to dissipate energy because the surface support
transfers some of the impact energy to the bolts. The energy of the
broken rock depends on its impact velocity and the volume or the
mass of broken rock. Because the mass of broken rock increases
with increasing bolt spacing, the demand on the surface support
capacity strongly depends on the bolt spacing.
For example, for a bolt pattern at 1 m ´ 1 m, a 0.9 m deep (steep
61°) wedge weighs 1.2 t/m and it impacts the retention system
with an energy of 0.6, 2.3, 5.3, and 9.4 kJ/m2 for 1, 2, 3, and 4 m/s
initial velocity, respectively. For a 1.4 m ´ 1.4 m pattern with a
1.2 m deep (3.2 t/m) wedge, the corresponding average energy
values are 0.8, 3.3, 7.4 and 13.1 kJ/m2. These values are indicated
in Figure 5-15, by red dashed and blue arrows on the energy–
deformation chart for mesh-reinforced shotcrete (Kaiser et al.

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
1996).
35
#6 mesh

30 severe (thin)
severe (thick)
moderate
25
Kinetic energy (kJ/m2)

minor
pull test
20
pull test

15
mesh-reinforced
shotcrete
10

5
#6 gauge mesh

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Deflection (m)
Figure 5-15 Anticipated deflections of mesh-reinforced shotcrete for two ejection scenar-
ios: (1) Red arrows from bottom to top: ejection of 1 m ´ 1 m ´ 0.9 m deep wedge at 1, 2,
3, and 4 m/s and (2) blue arrows for ejection of 1.4 m ´ 1.4 m ´ 1.2 m wedge (modified
energy graph from Kaiser et al. (1996))
This example illustrates that the central deflection increases as the
bolt spacing is widened (compare blue with red arrows). The
mesh-reinforced shotcrete will get excessively deformed (> 0.15
m central deflection) and moderately damaged when impact veloc-
ities surpass 4 m/s and the impact volume exceeds 1 m3 or 2.7 t.
Tight bolting patterns are needed to ensure that the retention sys-
tem can retain its integrity.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


204 Rock support system capacity

Load transfer capacity of connections


The retention system has to transfer the load to the bolts or vice
versa and it has to withstand the impact of point loads exerted by
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

the bolt plates. The puncture capacity therefore exceeds the high-
est bolt load. In other words, in addition to designing a retention
system for energy dissipation by impacting rock and displacement
compatibility to maintain the support system’s integrity, it must be
designed to prevent puncture failure due to impact loading by the
bolt plates (from the excavation side).
Because the impact is equal to the product of mass m times the
�⃗, the force exerted on the bolts by a single wedge
acceleration 𝑎𝑎
theoretically is m𝑎𝑎�⃗ (i.e., if a wedge is impacting in four adjoining
sections around a bolt). For strainbursts, the acceleration 𝑎𝑎 �⃗ is pro-
portional to the rupture time, which is normally unknown. How-
ever, the maximum possible force is the peak load capacity of the
bolts or cables making up the support system. For this reason, the
‘MAX yield force on the retention system’ is listed in the tables of
Figure 5-12 to Figure 5-14. This is the point load that a surface
support has to survive.
For the example with Split set, rebar, D-bolt and cables, presented
in Figure 5-12, the maximum point loads from individual support
components range from 50 to 225 kN. The retention system and
the connections must be designed to survive these point loads.
Larger plates and straps under plates assist in raising the puncture
capacity and thus are needed for those bolt types with high ‘MAX
yield forces’. Interestingly, Split sets with the lowest bolt load are
frequently furnished with the largest plates.
Even though there are only a few dynamic test results of surface
support component with rockbolt force monitoring data available
(other than from the test conducted using the Swiss facility), it
must be expected that the dynamic bolt forces, before failure of
surface support elements suitable for burst-prone ground, exceed
100 kN.
Static pull tests published in the literature (incl. Kaiser et al. 1996)
suggest that standard mesh or thin mesh-reinforced shotcrete with
nominal thickness of ≤ 100 mm fails at bolt forces between < 30
and 40 kN. Tests by Bucher et al. (2013) indicate that the maxi-
mum puncture capacities are about 50 and 65 kN for soft and stiff
assemblies, respectively. Player et al. (2008) tested 5.6 mm diame-
ter standard 100 mm square weld mesh and compared it with 4
mm diameter high strength chain-link mesh provided by Geobrugg
(at 1.3 m bolt spacing). The rupture loads were 45 to 60 kN for the
weld mesh at about 200 mm central deflection and 85 to 170 kN
for the high strength chain-link mesh at 300 mm central deflection.
The 5.6 mm weld mesh, tested by Player et al. (2008), would
therefore not be adequate for the above quoted base design and the
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 205

4 mm high strength chain-link mesh would only be adequate if


large plates were used to prevent failure (at 85 to 170 kN). Even if
failure was prevented, the mesh would, in both cases, deform
heavily with 200 to 300 mm or more central deflection.
Because an 8 mm weld mesh has an almost 2-times larger steel
cross-section, it can be assumed that such a heavy mesh should
resist point loads of 90 to 120 kN. It might survive the point loads
imposed by the bolts of the base design but only if very large
plates were used. Because 8 mm mesh is heavy and difficult to
install, 8 mm mesh straps instead of full mesh sheets are often
used as a means to improve the retention capacity of mesh. From
this discussion, it follows that, at a minimum, moderate mesh with
heavy straps and large plates are required to prevent puncture fail-
ure.
It also follows that special measures are required to prevent failure
at the bolt/mesh connection even if high strength mesh is used. For
this reason, the use of heavy straps or mesh embedded in shotcrete

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
is highly recommended for rockburst support systems. It has been
proven to be very beneficial in ensuring the integrity of surface
support systems.
Kaiser et al. (1996) showed that shotcrete more than doubled the
load bearing capacity of standard mesh at less than 100 mm cen-
tral deflections (Figure 5-16). Accordingly, a puncture load capac-
ity of 90 to > 120 kN for moderate to heavy weld mesh embedded
in shotcrete can be expected. Tests by Kirsten and Labrum (1990)
confirm this for uniform distributed loading but show that the load
capacity rapidly drops for point loading and for fibre-reinforced
shotcrete (Figure 4-19). As a matter of fact, for point loading the
mesh-reinforced shotcrete reverts basically to the capacity of the
mesh alone (20 to 60 kN). It follows that standard mesh or shot-
crete with imbedded standard mesh cannot survive point loads im-
posed by dynamically loaded bolts and remedial measures such as
heavy straps or mesh over shotcrete (double-layer systems) are
required to ensure that the connections between bolts and surface
support are not inferior.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


206 Rock support system capacity
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Figure 5-16 Shotcrete-to-mesh load and energy ratios as a function of displacement


(Kaiser et al. 1996).
Based on the rupture loads obtained by Player et al. (2008) and the
assumption that mesh imbedded in shotcrete can carry 2-times the
capacity of the mesh, the following limiting capacities are to be
expected:
- 5.6 mm mesh in shotcrete: 90 to 120 kN (confirmed by Kirsten
and Labrum (1990)); (Brändle et al. 2017)
- 8 mm mesh in shotcrete: 180 to 240 kN; and
- high strength mesh in shotcrete: 190 to 340 kN7.
It is for these reasons that failure of connections between tendons
and the surface support is less likely when shotcrete is used. The
benefit of the shotcrete is to distribute the point load and act as if
loaded through a large plate. However, it also follows that fibre-
shotcrete and mesh-reinforced shotcrete with light to moderate
mesh are inadequate to transfer anticipated loads to the tendons.
Heavy mesh, straps-over-shotcrete or a two-layer system with
mesh-over-shotcrete is needed to provide an adequate puncture
capacity to sustain point loads imposed by the bolts ranging from
180 to 240 kN for the above quoted base design.
- El Teniente mine’s double-layer system (Figure 4-24) with high
strength chain-link mesh over mesh-reinforced shotcrete held
with heavy rebar and cablebolts satisfies this criterion.
- Heavy straps over mesh-reinforced shotcrete would likely also
satisfy this criterion but it exposes workers to the risk of ‘shot-
crete rain’ (spalling of shotcrete over mesh). Hence, mesh over
mesh-reinforced shotcrete is preferred.

7
Brändle et al. (2017) measured impact forces exceeding 400 kN and static puncture
loads of 180 kN.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 207

These examples illustrate that a high capacity retention system is


required to transfer loads from the retention system to the bolts or
to prevent bolts from damaging or puncturing the retention system.
For these reasons, the E100 and E200 range values listed in Table 4-
4 are considered to provide practical thresholds of retention sys-
tem capacities.

5.3.5 Combined direct impact loading capacity of integrated


support systems
As indicated in Section 5.2.1 and earlier in Section 5.1, for direct
impact loading conditions, the approximated support system ca-
pacity – bolting plus surface capacity – can be obtained by adding
between 0- and 1-times the surface support system capacities
listed in Table 4-4. The worst-case scenario (aS = 0) corresponds
to very stiff surface support systems that immediately transfer the
entire load to the bolting system. The best-case scenario (aS » 1)
represents a relatively soft retention system that dissipates the en-

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
ergy that equals its energy capacity before it transfers the remain-
ing energy to the bolting system. Unfortunately, the multipliers
obtained from impact tests (Section 5.2.1) cannot be adopted for
this purpose as the bolting systems at the time of impact in the
field rarely correspond with those tested using the Swiss or South
African test frames. However, because it is likely that some ener-
gy is dissipated by the retention system, it is reasonable to assume
that the integrated system capacity is:
Esystem = SEbolt plate + (0.5 to 0.9) Esurface, (5-3)
and because a central deflection between 100 and 200 mm is a
reasonable design goal, with
Esurface = ½ (E100 + E200) (5-4)
from Table 4-4, the integrated energy capacity can be approximat-
ed by:
Esystem = SEbolt plate + (0.25 to 0.45) (E100 + E200). (5-5)
This capacity sharing model is implemented in the ‘support sys-
tem characteristics’ spreadsheet (Support system characteristics
v1803.xlsx).
The respective surface capacities are superimposed in Figure 5-12
to Figure 5-14 on the total and remnant bolting system capacities.
For this purpose, it is assumed that the surface support capacity is
gradually mobilized from the point of installation at dbolt plate = 0 to
a chosen mobilization threshold (specified by the user) and then
gradually lost to zero at twice this displacement threshold. It is
also assumed that measures have been taken to ensure that the
point load capacity is sufficient to prevent puncture failure.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


208 Rock support system capacity

For example, for a retention system with a moderate mesh and


mesh straps:
Esystem = SEbolt plate + (0.25 to 0.45) (13 to 21)
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

= SEbolt plate + (3 to 10) kJ/m2. (5-6)


For this relatively soft surface support, the upper end of the range
is meaningful and 10 kJ/m2 can be added to the bolting system ca-
pacity if combined with a moderate mesh and mesh straps (and
weakest link issues have been prevented).
The results for the baseline bolting system with mesh-reinforced
shotcrete and threshold values of 100 and 150 mm are shown in
Figure 5-12b and Figure 5-17, respectively.
Split Set,Rebar,D-bolt at plate,Single Cable 1,
50
Mesh-reinforced reinforce SC Cummulative bolting capacity
45
Support system capacity
40 Remnant bolting system
capacity
35 Remnant support system
capacity
Energy capacity (kJ/m2)

30

25

20

15

10

0
0 50 100 150 200 250
Displacement on bolt system (mm)

Figure 5-17 Energy capacities of the baseline bolting system with Split sets, rebar, D-
bolts and single cables with mesh and straps for a central deflection threshold of 150 mm.
The remnant energy capacity after 100 mm is shown in red.
Figure 5-17 illustrates that the baseline design has a substantial
energy dissipation capacity of 28.7 kJ/m2 if none of the bolts and
the surface support are pre-deformed (black dotted curve). How-
ever, at 50 mm wall/plate displacement almost 50% of the ulti-
mate capacity, about 9 kJ/m2 of the bolting system and about 4
kJ/m2 of the surface support capacity, would already be consumed.
After 100 mm, the remnant bolting system capacity is next to zero.
The remnant capacity if the integrated support system of about 11
kJ/m2 would theoretically come from the surface support, but only
if the bolting system were stable, which is unlikely because most
its load capacity has been consumed at dB = 90 mm (see Figure
5-12a). As a consequence, for the baseline design, only approxi-
mately 24 kJ/m2 at dB = 90 mm can be relied upon. As explained
above, of this merely about 10 kJ/m2 (24 minus 14 kJ/m2) would
remain after 50 mm of pre-deformation.
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 209

Furthermore, it is unlikely that this system would be able to pro-


vide a 225 kN puncture resistance (needed for the cables) at this
deformation stage. Hence, the integrated support system will like-
ly only offer between 20 and 25 kJ/m2 energy dissipation capacity.
For this baseline design, the ultimate central deflection for the in-
tegrated system at the point of failure will be 300 mm (dB = 150
mm plus the central deflection of surface support system dS = 150
mm) or 200 mm after 50 mm of pre-deformation. However, if a
rockburst were to occur after a bolt displacement of 90 mm, with a
mean displacement of the surface support system of 90 mm for a
total central deflection of 180 mm, this baseline support would at
best provide an energy capacity of 10 kJ/m2.

5.3.6 Maximum practical energy support limit (MPESL)


Kaiser et al. (1996) stated that there are practical and economic
limitations that dictate maximum practical support dimensions
such as bolt spacing or shotcrete thickness. They concluded that it

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
is not practical to provide more than 50 kJ/m2 of energy absorp-
tion capacity with the then-available support technology, i.e., the
maximum practical support limit (MPSL) in terms of energy ca-
pacity (more appropriately termed maximum practical energy lim-
it (MPEL)) of a virgin support system is approximately 50 kJ/m2.
This view is not shared by Ortlepp and Swart (2002). In their
opinion, based on their test data, the MPEL can be much higher
than 50 kJ/m2. This may be true for ideal loading conditions as
discussed in the previous subsections and in Chapter 4. With the
development of the next generation of yielding bolts (e.g. MCB33
conebolt, D-bolt, etc.) and the high strength wire mesh (e.g., Tec-
co mesh, dynamic mesh, etc.), a system including bolts, mesh,
straps, and rocks, may be able to absorb a larger amount of energy
but only if all support components are simultaneously mobilized
and loaded in such a manner that they reach their individual ener-
gy capacities, and rock is retained to dissipate energy by friction.
The double-layer system developed by El Teniente mine is quoted
as having a cumulative energy capacity of 88 kJ/m2 by Joo (2017)
(50 kJ/m2 from 25 mm rebar at 150 mm, 18 and 12 kJ/m2 from
two layers of mesh at about 250 to 350 mm, 7.75 kJ/m2 from cable
at 80 mm, and 1 kJ/m2 from shotcrete). For a 1.2 m ´ 1.2 m bolt
and 2 m ´ 2 m cable pattern, the cumulative energy capacity is 70
kJ/m2 (35 kJ/m2 for rebar + 4 kJ/m2 for cable + 31 kJ/m2 for mesh
and shotcrete). When tested at the Swiss test facility (Figure 4-15),
this double-layer system of shotcrete and mesh with 150 mm2 ca-
bles and 490 mm2 (25 mm diameter) rebar deflected between 0.2
and 0.6 m and generated maximum anchor loads between 155 and
429 kN when impacted by 30 to 60 kJ. This is equivalent to be-
tween 21 and 42 kJ/m2 for a 1.44 m2 area between rebar or 12 and
Draft manuscript – Copyright protected – Cai and Kaiser 2018
210 Rock support system capacity

24 kJ/m2 for an equivalent area of 2.5 m2 between rebar and cables.


As discussed in Chapter 4 (Section 4.6.3), the energy dissipated at
a MPDL of 0.3 m by this double-layer system, including rockbolts,
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

cables and surface support, was at best between 15 and 24 kJ/m2


without debonding and between 30 to 45 kJ/m2 with debonding of
rebar and cables.
It can be concluded that when a freshly installed support system is
loaded by more than 50 kJ/m2, the associated displacements typi-
cally become excessive (> MPDL), which can cause operational
problems or disrupt the rock mass and the support system’s integ-
rity. Therefore, the authors reiterate that the maximum practical
energy limit – MPEL, previously called MPSL, is 50 kJ/m2 and
that strategic measures must be used to reduce the rockburst haz-
ard if this limit is exceeded (Chapter 3). Avoiding conditions with
high energy release potential is the most prudent and often the
most economical approach when the MPSL is exceeded.

5.3.7 Indirect impact loading capacity of bolting systems


As illustrated by Figure 5-4, there are two loading mechanisms
that need to be considered when assessing the support capacity –
direct loading near the excavation surface (near the plate) and in-
direct loading inside the burst volume at some distance from the
wall. Furthermore, as indicated in Chapter 4, whereas the load ca-
pacities are comparable for these two loading mechanisms, the
displacement and energy capacities of many bolt types are much
higher (potentially double) if the bolts are indirectly loaded.
If bolts are uniformly strained due to a constant rock mass bulking
distribution in the burst volume (red full line in Figure 5-4), the
displacement capacity of the bolts will rarely be exceeded as duc-
tile steel can accommodate in excess of 10% strain or > 100 mm
length change per metre (even rebar can yield > 9% or > 90
mm/m). More brittle steel, such as the steel used in high strength
cables, can only survive a minimum of 35 mm/m.
During strainbursts, the indirect capacity is mobilized inside the
burst volume and the displacement profile depends on whether
and where displacement localization occurs (red dashed line in
Figure 5-4). High displacement gradient and thus strain localiza-
tion must be expected at the transition from the burst volume to
the stable ground (behind the burst volume). Such strain localiza-
tions may drastically reduce the overall displacement capacity and
lead to bolt failures at this interface. For example, for a rebar with
a stretch length of 100 to 200 mm on each side of a localization,
the displacement capacity during indirect loading is only 18 to 36
mm per localization. For a high strength cable, it may be as low as
10 to 30 mm per localization (depending on the stretch length). On
the other hand, if such localizations were evenly distributed and

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 211

separated by at least twice the stretch length, these bolts could ac-
commodate displacements in the order of 90 mm/m and 35 mm/m
for rebar and high strength cables, respectively. Therefore, the
displacement capacity of these bolt types can be highly variable
and it must be assumed that the internal capacity could be reached
if local relative displacements were to exceed 10 to 40 mm/m. As
a consequence, the internal energy dissipation capacity is variable
with capacities potentially dropping to as little as 20 to 40% of the
peak values if localization occurs.
For yielding bolts with ductile steel (e.g., 15 to 20% elongation for
D-bolt) and paddle spacing of 0.65 to 1 m, the displacement ca-
pacity ranges from 100 to 200 mm/m. This leads to superior ener-
gy dissipation capacities, theoretically even if localization occurs.
Such high displacement capacities can rarely be reached because
of support component incompatibilities. If this displacement is lo-
calized, the rock mass will open up facilitating shear, which in
turn reduces the bolts’ axial load capacity (see Section 4.2.2).

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Whereas the internal energy dissipation capacity may be higher
than for non-yielding bolts, in reality, it may not be as high as the
split tube test results suggest.
Conebolts rely on cone plow to provide the energy dissipation ca-
pacity and offer very high displacement capacities that exceed
those of other bolt types. However, because of associated rock
mass disruption and bulking (facilitating shear), the high dis-
placement capacities obtained from pure tensile loading tests can
again rarely be reached. The ultimate internal energy dissipation
capacity is theoretically very high but, as for all bolt types, in
practice, it is not as high as split tube test results suggest.
The practical implication is that the internal displacement and en-
ergy dissipation capacities of bolting systems can be much higher
than the direct loading capacities, as long as strain localization in-
side the rock mass is prevented. Uniform displacement patterns
are best achieved by combining relatively stiff bolts, e.g., rebar
that help minimize localizations, with yielding bolts that provide
superior energy dissipation capacity at larger displacements.
These considerations have to be respected in order to arrive at in-
ternal displacement and energy capacities for support design con-
sidering energy dissipation by the rock mass reinforcement Erein-
forcement.

Field evidence suggests that a support system either fails by direct


loading with unravelling of the broken rock at or between tendons,
or by internal reinforcement failures if the direct loading capacity
is not exceeded. In other words, the indirect loading capacity can
be mobilized if direct loading failure is prevented. Hence, the di-
rect and indirect energy capacities cannot be mobilized simultane-
ously and the respective capacities cannot necessarily be superim-

Draft manuscript – Copyright protected – Cai and Kaiser 2018


212 Rock support system capacity

posed. The term Ebolt = S(Ebolt plate + Ereinforcement ) in Eq. (5-2) is on-
ly valid if the energy is dissipated by direct loading before the en-
ergy is consumed by internal loading, or vice versa, if energy is
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

dissipated internally before the surface support is loaded.


During a strainburst, the reinforcement inside the burst volume
will be strained (dissipating some indirect energy) and the impact
energy on the burden or surface support will be reduced. In this
case, indirect and direct displacement and energy capacities will
be consumed but rarely will the maximum indirect and direct ca-
pacities be reached simultaneously. Furthermore, if localization
occurs, the surface support may not be deformed as all loads are
directly imposed through the ‘burden’ onto the bolt plates.

Figure 5-18 Scenarios 1 to 4 for integrated system capacity assessment.


From these considerations, it follows that there are several scenar-
ios (Figure 5-18) to be assessed in a support design process:
- Scenario A: The bulking displacement is directly imposed on the
bolts. The direct loading scenario without energy dissipation by
the surface support applies: Esystem = S(Ebolt plate).
- Scenario B: The bulking displacement is imposed on the surface
support and to the bolts. The direct loading scenario with energy
dissipation by the surface support shared with bolts applies: Esys-
tem = SEbolt plate + Esurface. This load model also applies to energy
transfer from remote seismic sources.
- The bulking displacement is imposed on the reinforcement (due
to strainbursting inside the reinforced rock mass). Some of the
indirect energy dissipation capacity will be consumed before
Scenarios A or B are activated. Unfortunately, failure may occur
when either the indirect or the direct loading capacity is reached.
If strain localization is prevented, the indirect loading capacity of
a grouted bolt can be about twice the direct bolt loading capacity
and Esystem = S((1 to 3) Ebolt plate) + Esurface. However, if localiza-
tion is not prevented, the indirect bolt loading capacity will be
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 213

comparable to the direct bolting capacity and Esystem = S((1 to 2)


Ebolt plate) + Esurface. Because it is difficult, if not impossible, to
prevent localization during rockbursts, it is reasonable to assume
that the second condition (where localization is not prevented)
should form the basis for a design. The multiplier 2 applies if the
bolting system inside the burst volume is equal in capacity to the
one near the wall [Scenario C].
- Scenario D: The bulking displacement is imposed on the gabion
or the reinforced rock mass arch (due to strainbursting behind the
rock mass reinforcement). For this case, the direct loading model
is not applicable because the reinforced support arch (gabion) is
impact loaded and has to move as a package. The entire gabion
becomes the retention system and it will dissipate energy as it is
moved into the excavation. The cables reaching beyond the ga-
bion will contribute to the energy dissipation by indirect loading.
This load model also applies to situations with energy transfer
from remote seismic sources if the frequency of the incoming

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
waves is sufficiently low to accelerate the entire gabion. This
mechanism will be explored in Section 5.4.6.
In summary, if strainbursting is expected:
- inside the reinforced rock mass but very close to the wall, the in-
tegrated system capacity is Esystem = S(Ebolt plate) + Esurface;
- inside the reinforced rock mass but far enough from the wall, the
integrated system capacity is Esystem = S(1 to 2) Ebolt plate;
- far from the wall behind the reinforced rock mass with some
deep-seated anchorage, the integrated system capacity is the di-
rect loading energy capacity of the anchorage plus the energy
consumed by the reinforced rock mass (gabion). The latter de-
pends on the failure mechanism of the reinforced rock arch. For
a flat wall, it can be approximated by the energy consumed by
sliding the gabion into the excavation (this scenario is explored
in Section 5.4.6). If stress-arching occurs in curved reinforced
rock arches, energy will also be consumed by the gabion.
It is difficult to anticipate the exact strainburst scenario and it is
reasonable to assume that the integrated support system capacity is
well represented (potentially underestimated) by the direct capaci-
ty model introduced in this chapter: Esystem = S(Ebolt plate) + Esurface.
Figure 5-19 illustrates the consequences of these scenarios (in
counter clock wise manner):
- Top left: Scenario D with burst kicking out lower part of the wall
causing indirect loading;
- Mid left: Scenario D combined with B (or A) causing indirect
loading with local direct loading failure;

Draft manuscript – Copyright protected – Cai and Kaiser 2018


214 Rock support system capacity

- Bottom: Scenarios A to C without surface support failure; on the


right side, Scenario B almost caused surface support failure (and
there may have some Scenario D process in play; see next)
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

- Middle and top right: Scenario D failures with burst kicking out
upper part of the wall causing indirect loading.

Figure 5-19 Gabion failure modes: A and B causing gabion surface support failure and D
causing gabion rotation.
The cartoon in the middle of Figure 5-19 highlights the associated
rotational failure modes. It is also possible that the gabion trans-
lates as shown by Figure 5-20 with associated floor heave (or
‘ploughing of muck’).

Figure 5-20 Gabion failure modes: D causing gabion translation with ploughing of floor.
With respect to puncture capacity, the values listed in Table 4-4
are applicable for Scenarios A and B. For Scenarios C and D, the
retention system is impacted via broken rock between the burst
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 215

location and the bolt plates. This rock distributes load and drasti-
cally increases the puncture load capacity of the integrated system
(gabion). It is for this reason that a well-designed gabion with tight
reinforcement patterns provides superior retention system behav-
iour. The images in Figure 5-19 show that no puncture failure oc-
curs when the reinforced rock mass deforms in unison, i.e., de-
forms as a gabion.

5.4 Rock support systems – ground-truthing


The true capacity of a rock support system can only be evaluated
when loaded by actual rockbursts. None of the above simulated
tests or experiments can capture the complex interaction between
support components and between rock and rock support. Ground-
truthing is the ultimate experiment to identify strengths and weak-
nesses of rock support systems. The disadvantage of this approach
is that the test time cannot be controlled and the damage evalua-

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
tion is largely based on visual inspection rather than quantitative
observations.
When ground-truthing support systems, it is critically important to
identify the predominant rockburst damage process, whether it is
strainbursting, a shakedown or an energy-driven failure caused by
remote seismic events. Each process will impact a support system
in a different manner. Whereas a given support system may resist
loading from one mechanism, it may not survive loading by an-
other mechanism. As explained throughout this book, damage
caused by dynamic excavation failures must be differentiated from
damage caused or dominated by dynamic disturbances from re-
mote seismicity. This distinction was rarely made in the past as
discussed in the following case examples illustrating the nature of
an approach to ground-truthing. The case histories are all qualita-
tive in nature as no local instrumentation other than seismic moni-
toring systems were in place.

5.4.1 Big Bell Mine, Australia


Five seismic events, measuring ML = 1.9, 2.2, 1.7, 2.4, and 2.2,
occurred at Big Bell Mine in Australia and affected the 485- and
460-footwall drives in August and November of 1999, and the
535-footwall drive in June of 2000 (Turner and Player 2000). The
support system installed at the mine site up to late 1999 included
Split sets, end-anchored rockbolts (replaced with steel pipe tubular
bolts (TGBs) in January 1999), mesh (RF81, diameter 7.6 mm,
yield strength 500 MPa), with cablebolts at intersections. The
rockburst damage to the rock support systems typically included
failure of Split Set rings, pulling out of Split sets, failure of end-
anchored bolts, TGBs, and separation of mesh from the hanging

Draft manuscript – Copyright protected – Cai and Kaiser 2018


216 Rock support system capacity

wall shoulder at mesh overlaps. An upgrading of the support sys-


tem included grouting of cablebolts inside Split sets. These
stronger, but relatively stiff holding elements were not effective in
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

preventing rockburst damage. A major rockburst of ML = 2.4 oc-


curred on the 485 Level on November 25, 1999, causing 40 m3 of
fallen rock. The failure mode illustrated by Figure 5-21 was iden-
tified as a seismically induced fall of ground (Turner and Player
2000). According to Diederichs (2014), the damage causing fail-
ure mechanism may hypothetically be classified as a structurally
controlled strainburst with consequential unravelling or seismic
shakedown.

Figure 5-21 Rockburst damage from the November 25, 1999 event at the 485 footwall
drive at Big Bell Mine (Turner and Player 2000).
The rockburst locally destroyed all support elements in areas with
a calculated, cumulative static support capacity of 230 kN/m2
(Turner and Player 2000), assuming simultaneous activation of all
support elements8. This drift section included Split sets at about 50
kN/m2, twin strand cablebolts at 125 kN/m2, end anchored bolts at
25 kN/m2, and mesh at 30 kN/m2. The total system capacity is not
simply a summation of all support element installed. As discussed
in Chapter 3, the weakest link often dictates the system’s capacity.
Whereas the authors did not examine the condition at the site, the
image in Figure 5-21 suggests that it is reasonable to assume that
the actual effective load capacity was smaller (at best 125 kN/m2
from the cables) because:
- The Split sets and end-anchored bolts were too short, at least
in the centre, and the retention system failed due to the loss of
anchor capacity (leading to a large bolt spacing at the time of
failure).

8
Assuming that all support elements are available to act at the same time is an unrealistic
assumption considering the installation sequence and non-compatibility of various in-
stalled support elements.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 217

- The fragmented muck pile supports the assumption that an un-


ravelling of the stress-fractured rock between the cablebolts
must have occurred.
- The cablebolt capacity may also have been compromised by
the associated de-clamping and the cablebolt’s load capacity
was exceeded as indicated by the snapped cables at the plate.
A decision was made to introduce a rockburst resistant support
system comprised of South African conebolts (approximately 1 m
´ 1 m pattern), M71 or M61 weld mesh (7.1, 6.3 mm in diameter,
100 mm ´ 100 mm, 450 MPa yield strength; see Appendix E),
rubber padding (between mesh and surface plates) and twin dome
plates (Player 2004). The mesh was changed in late 1999 from
RF81 to M61 (with additional strength welds specifically designed
for mining applications). A seismic event (ML = 1.7) occurred on
November 25, 1999, in the 460-footwall drive and damage was
prevented or contained by this support system. This convinced the
mine to accept conebolt-based support systems to prevent seismi-

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
cally induced shakedowns.
In retrospect, it is of interest to note that the role of the conebolt-
based support systems was not primarily to dissipate energy but to
prevent unravelling by ensuring a coherent and deformable back.
In this manner, the weak links were eliminated and the compatibil-
ity between all support components was ensured.
Further seismic events occurred on April 6, 2000 near the 510-
footwall drive and on May 23, 2000 near the 535-footwall drive.
The installed rockburst support system again prevented or mitigat-
ed excavation damage (Turner and Player 2000).

5.4.2 Brunswick Mine, Canada


In the late 1990s Brunswick Mine experienced a number of rock-
bursts with increasing severity. Within the framework of this book,
and in retro perspective, it is now evident that two mechanisms
were at play. At intersections, the seismicity caused seismically
induced shakedowns (this was evident at many locations at
Brunswick Mine). In drifts the damage was predominantly caused
by triggered or dynamically loaded strainbursts with slabbing,
buckling or bulking causing rock and support (e.g., shotcrete)
ejection.
Many intersections collapsed despite the fact that they were sup-
ported by rebar, cablebolts, mesh screen, and in some places with
shotcrete. It is now evident that the retaining function failed
(weakest link issue) and facilitated unravelling at the intersections.
This is illustrated by the photo of an unravelled intersection in
Figure 5-22a.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


218 Rock support system capacity

At the same mine, drift walls, supported with standard bolting and
shotcrete (Figure 5-22b) failed in a violent manner by triggered or
dynamically loaded strainbursts whereby a shallow shell of rock,
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

just behind the shotcrete, failed, ejecting rock and shotcrete be-
tween widely spaced bolts (note: rock failed around many bolts
with intact plates). This image shows again that the retention sys-
tem was inadequate and poorly connected to the bolts, illustrating
the weakest link issue.

(a)

(b)
Figure 5-22 Failure modes observed at Brunswick Mine: (a) unravelling due to seismic
shakedown; and (b) dynamically loaded strainburst (photos by P.K. Kaiser, 1999).
This increase in violent excavation failures at Brunswick Mine led
to the urgent development of an integrated, yielding support sys-
tem (Simser et al. 2002a). The South African conebolts (Jager
1992) were modified to accommodate resin-grouting practices in
Canada. The original intent was to proceed with a fully instru-
mented field trial of the modified conebolts (MCB) using simulat-
ed rockburst techniques. However, the urgency of the situation at
the mine site dictated an immediate underground installation with-
out verification testing.
Greased debonded MCB38 conebolts at 1 m spacing with 150 mm
´ 150 mm, 9.5 mm thick domed plates were used with #0-gauge
mesh straps (Appendix E). These mesh straps were installed to
overlay the diamond mesh (5 cm aperture, 5 mm diameter) as the
second-pass system on top of the standard support system. It com-
prised 2.3 m long rebar on a 1.5 m ´ 1.5 m pattern with shotcrete

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 219

and #9-gauge (3.66 mm diameter) weld mesh (Simser et al.


2002b). Regular shotcrete was applied over the weld mesh; steel
fibre reinforced shotcrete was used in the walls and in some cross-
cuts (e.g., 326-crosscut) no weld mesh screens were installed.
A close-up view of the (partially installed) rockburst support sys-
tem can be seen in Figure 5-23. It should be noted that by imple-
menting the much tighter bolt spacing, the retention and energy
dissipation capacity was increased.
A distinct boundary between stable and unstable ground is visible
in Figure 5-23a (red dashed line). This boundary is called the ‘ex-
treme edge’ at the mine site. Beyond the ‘extreme edge’ rock sup-
port system consisting of chain-link mesh, end anchored rockbolts
with a wider 1.5 m ´ 1.5 m spacing were installed. This support
was previously deformed by mining-induced deformations includ-
ing co-seismic deformations from seismic ‘hammering’.

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
#6-gauge chain-link
mesh + #0-gauge
mesh strap + MCB
conebolts (17.3 mm) Chain-link mesh
+ end anchored
rockbolts

(a)

#6-gauge chain-
link mesh + #0-
gauge mesh strap +
MCB conebolts
Rebar + shotcrete + (17.3 mm)
#9-gauge weld mesh

(b)
Figure 5-23 Ground-truthing at Brunswick Mine: (a) ‘extreme edge’ created by two types
of rock support systems; (b) drift roof and wall performance with two types of rock sup-
port systems (Photo courtesy: Brunswick Mine).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


220 Rock support system capacity

A series of rockbursts occurred between October 13 and 17, 2000,


and severely damaged the excavation at locations where standard
support was installed. The section that was supported by the rock-
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

burst support system suffered less or no damage. On Oct. 13,


2000, a ML = 2.0 rockburst event caused a shakedown collapse of
an intersection involving approximately 1900 t of material. A
small portion of the 326 cross-cut (a close-up view can be seen in
Figure 5-23b) immediately east of the intersection was supported
with MCB38 conebolts.
The right-hand side of this drift was supported by 2.3 m long
MCB38 conebolts in a 1 m × 1 m pattern with #0-gauge mesh
straps and chain-link mesh. The left-hand side of the drift was still
reinforced by a standard support system. Hence, the rockburst
support system was not yet fully installed at the time of the rock-
burst as the chain-link mesh was only suspended with 0.9 m end-
anchored bolts. The area in the foreground was supported by 2.3
m long rebar and steel fiber-reinforced shotcrete. After a rockburst
event with ML = 2.0, the right-hand side of the drift suffered no
visible damage whereas the left-hand side suffered support dam-
age with local bagging of mesh due to rock bulking and unravel-
ling of stress-fractured rock.
The intersection ahead of this drift collapsed despite the fact that it
was supported by rebar, cablebolts, mesh screen, and shotcrete. As
discussed above, this can be attributed to a seismic shakedown of
poorly retained ground. There was some speculation that the posi-
tive performance of the MCB38 conebolt supported drift section
was to be attributed to locally stronger ground or less dynamic
loading. However, close onsite examination after the event re-
vealed that the conebolts had fulfilled their yielding support role
as some conebolts were displaced as much as 180 mm (Simser et
al. 2002a). The reason why the left-hand side after the event still
contained the damaged rocks may in part be attributed to the fact
that the stable right-hand side of the drift reduced the effective
span for the left half. This may have reduced the unravelling po-
tential as reflected in the limited amount of broken rock in the
mesh. Consequently, only minor damage was observed in the rock
on the left-hand side.
Subsequent application of the MCB38-based rock support system
with thigh bolt spacing (1 m × 1 m) has served the mine well. It
proved to be successful in containing moderately severe damage
associated with dynamically loaded strainbursts resulting from
seismic events of up to ML = 2.7. This example illustrates not only
the benefit of deformable support but also the benefit of proactive
support maintenance. By adding support that has not been de-
formed by previous mining, the remnant capacity of the standard
support system has been effectively enhanced.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 221

Whereas this type of ‘field experiment’ with two support systems


in a single drift was not planned, the comparison of rockburst
damage on the left and right side of the drift demonstrates that
more deformable support systems can effectively limit rockburst
damage to underground openings.

5.4.3 Copper Cliff North Mine, Canada


On Sept. 11, 2008, a ML = 3.3 seismic event (at the 3050 Level)
occurred immediately after a crown blast in the middle 100-
orebody between 3050 and 3200 Levels at Vale’s Copper Cliff
North Mine in Sudbury, Ontario, Canada. This large rockburst
event, in association with other significant seismic events (with
ML > 0.7), caused widespread damage extending from 2700 to
3710 Levels around the 100/900-orebody region. In total, more
than 2500 t of material was displaced at various locations on dif-
ferent levels, and most of the damage was associated with geolog-
ical structures (Yao et al. 2009). The closest damage was observed

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
at a distance from the seismic source between about R = 50 and
250 m from the ML = 3.3 event. The maximum anticipated ground
motions, based on scaling parameters established for Creighton
mine in Canada, were in the order of PGV = 0.3 to > 1 m/s and the
maximum pseudo-dynamic stress pulse at the excavation walls
could have been in the order of 10 to > 20 MPa.
One of many reasons for the extensive damage was that the in-
stalled ground support was relatively light (a mix of resin rebar
and mechanical bolts with #6-gauge wire mesh (4.88 mm wire di-
ameter; see Appendix E)) with a rather limited holding and energy
absorption capacity. Furthermore, it was pre-deformed by previ-
ous mining activities and its remnant capacity was clearly less
than the maximum capacity of the installed support.
After the occurrence of the major seismic event, a mitigation plan
was put in place to ensure that the remaining orebodies could be
mined safely. The mine management adopted a modified cone-
bolt-based dynamic rock support system on top of the primary
support consisting of:
1) #4-gauge welded wire mesh (5.89 mm wire diameter), 1.98 m
FS46 (46 mm) friction bolts on a 1.22 m × 0.76 m pattern for wall
and 2.44 m resin rebar on a 1.22 m × 0.76 m pattern for back
(roughly 2 bolts per m2).
The secondary (rockburst) support overlay consisted of:
2) 2.34 m long modified conebolts on a 1.22 m × 1.83 m pattern with
#0-gauge mesh straps (roughly one bolt per 2 m2).
It is important to note that this support system not only had a
higher displacement capacity but it was not preloaded. Following

Draft manuscript – Copyright protected – Cai and Kaiser 2018


222 Rock support system capacity

the approach introduced earlier for support system capacity esti-


mation, it is seen that the:
- primary support has a maximum bolting system capacity of
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

about 16 kJ/m2 and an integrated support system capacity of


about 22 kJ/m2 (would drop to 7 and 10 kJ/m2 at 100 mm, re-
spectively); and
- secondary support overlay adds displacement and energy capaci-
ties such that the maximum bolting system capacity is increased
to about 24 kJ/m2 and an integrated support system capacity to
about 30 kJ/m2 (would drop to 7 and 13 kJ/m2 at 100 mm, re-
spectively).
In other words, this newly installed support system derives about
two thirds of its capacity from the primary support. The secondary
support overlay adds about 40 to 50% energy dissipation capacity.
Because of the heavy-duty friction bolts, the primary system
would lose its holding capacity at about 150 mm. The benefit of
the MCB therefore is to further extend the displacement capacity
of the integrated support system. This example again illustrates the
benefit of restoring the remnant capacity to the full capacity of a
support system.
On Feb. 2009, a ML = 2.4 seismic event occurred in the same area,
causing further damage to the drifts. All areas with the new prima-
ry support system, including the modified conebolt overlay, were
undamaged (Yao et al. 2009). Based on this experience and a
rockburst risk assessment, the mine planned to systematically rein-
force areas with a high rockburst risk using a deformable overlay
of yielding bolts and mesh straps.
It is important to note that the remnant capacity of this integrated
support system can again be consumed by mining-induced defor-
mations, such that its bolting system capacity and integrated sup-
port system capacity after 100 mm of wall displacement would be
reduced to 7 and 13 kJ/m2, respectively. Despite the use of a yield-
ing support overlay, PSM work will be required in areas where
about 100 mm of the support’s displacement capacity has been
consumed. Moreover, the bolt density was increased by adding
deformable support components and the addition of mesh straps
enhanced the retaining capacity. The latter reduced the unravelling
potential and the former provided additional displacement and en-
ergy dissipation capacities.

5.4.4 Kidd Mine, Canada


On January 6, 2009, a magnitude ML = 3.3 seismic event, believed
to have occurred on the North G Fault, just below the 7000 Level,
caused extensive damage to four levels (6800, 6900, 7000 and
7100) at Kidd Creek Mine, Timmins, Ontario, Canada. The event
triggered at least 30 stations of the Canadian seismic network and
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 223

the closest damage at the mine site was observed at about R = 10


to 20 m and the furthest damage at about R = 150 m from the ML =
3.3 event. Fortunately, the event occurred at about 4:40 a.m., as
the nightshift at the mine was going off duty and was heading to
surface. No injuries or fatalities were reported. The anticipated
ground motions (using the Creighton mine scaling parameters)
were in the order of PGV = 0.5 to >> 1 m/s and the pseudo dy-
namic stress pulse at the excavation walls would have been in the
order of > 20 MPa. Most of the severe damage was restricted to
within 10 to 20 m of a major shear rupture. The observed damage
affected about 1500 m of developments (drifts) on seven levels
(Counter 2010). One drift supported by previously deformed
standard rock support with mesh and rebar was severely damaged
and much of the excavation was filled with broken rock (Figure
5-24a). Several pieces of mining equipment were buried due to
this seismic event (one buried machine can be seen in Figure
5-24b).

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
(a)

(b)
Figure 5-24 Ground-truthing at Kidd Mine: (a) Damage in 68-01 Drive (supported with
varying length of rebar); (b) damage in 70-82 cross-cut beyond area supported with
modified conebolts (Photo courtesy: Kidd Mine).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


224 Rock support system capacity

Underground inspection after the event revealed that in areas rein-


forced with a deformable support system (modified conebolt-
based support in this case) were in relatively good shape when
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

compared with damage in adjacent areas where standard rock sup-


port (rebar and meshes) was used (Figure 5-24b). There is no
question that deformable bolts are beneficial as they increase the
displacement and energy dissipation capacity. It has to be pointed
out that the ‘standard rock support’ was previously deformed and
therefore had a remnant capacity that was less than optimal. The
areas with the ‘deformable support system’ were newer and thus
must have experienced less preloading. As a consequence, it had a
higher remnant capacity.
These case histories reveal the performance of various rock sup-
port systems under rockburst conditions and demonstrate that
more deformable support systems help mitigate rockburst damage.
These examples, however, also demonstrate the great benefit of
proactive support maintenance (PSM) to ensure that a high rem-
nant capacity is available at the time of a rockburst.
Active PSM strategies are now being implemented in burst-prone
areas of the Grasberg mine (Freeport-McMoRan, Indonesia; see
Section 5.4.6) and Cadia Operations (Newcrest, Australia).

5.4.5 El Teniente TEN-SUB6, Chile


Codelco, a Chilean state-owned copper mining company, was a
member of the Canadian Rockburst Research Program and the ex-
perience from TEN-SUB6 (a sub block caving panel at El Ten-
iente mine) led to the ground-motion-centric support design ra-
tionale reflected in the Canadian Rockburst Support Handbook
(1996; CRBSHB). A recent underground visit to El Teniente mine,
some 20 years later, revealed why major adjustments presented in
this book had to be made to improve the support design approach
published in the CRBSHB. The detailed reasoning is presented in
Chapter 1. The main deficiency was that the contribution of stored
strain energy or the influence of variations in the deformation po-
tential (mine stiffness) was not considered in 1996. Today, it is
recognized that a sound support design has to respect the energy
released from the burst volume itself and the rock mass surround-
ing it (as it often dominates the energy demand on the support).
This recognition has led to the currently adopted support design
approach presented in this book.
Furthermore, the concept of support capacity consumption result-
ing from mining-induced displacements and from seismic ‘ham-
mering’ was not considered in 1996. Hence, three additional as-
pects have to be respected for support design purposes:
- spatial patterns of mining-induced displacement;

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 225

- spatial patterns of cumulative tangential dynamic stress or cumu-


lative co-seismic displacements; and
- spatial patterns of mine stiffness or deformation potential.
As described above, El Teniente mine has steadily increased the
energy dissipation capacity and with it its support displacement
capacity, culminating in the current double-layer design (see
Chapters 1 and 5). This advanced support system is discussed at
the end of Chapter 4 and is used in Section 5.3.6 to illustrate the
maximum practical energy support limit (MPESL).
El Teniente mine uses G80-4 high strength mesh for this purpose
and estimates that the total ultimate energy dissipation capacity of
the support system is 88 kJ/m2 with 50 kJ/m2 from the 25 mm di-
ameter threadbar, 18 and 12 kJ/m2 from the two mesh layers, 7.8
kJ/m2 from the double cables and 1 kJ/m2 from the shotcrete9. This
is a theoretical value as it is only achievable if all support compo-
nents are simultaneously deformed to reach their respective capac-
ities at the same pace and do not prematurely fail. In reality, this is

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
impossible to achieve. In the case of the double-layer support sys-
tem adopted at El Teniente mine, the displacement capacity of the
rockbolts will likely be exhausted before the full mesh capacity
can be mobilized.
If the cables were able to deform about 75 mm, the threadbars
about 150 mm, and the surface support could deflect about 200
mm, the support system model would predict the behaviour shown
in Figure 5-25. This is roughly equivalent of the capacity (88
kJ/m2) quoted by the El Teniente mine. This support system would
apparently also offer a good remnant capacity of about 54 kJ/m2
after 75 mm displacement.
However, the above quoted energy dissipation values for the dou-
ble-layer El Teniente mine support system were obtained from
split-tube test data and therefore are only valid for indirect loading.
For example, for a 25 mm threadbar to dissipate 50 kJ energy, a
debonded section greater than 1.2 m is needed in a split-tube dy-
namic loading test. In direct loading, the energy capacity of the
bolts may be as little as 50% or less (assume 0.6 m free stretch
length). If the threadbars are fully grouted, the energy capacity for
direct loading can be even lower (< 10 kJ). The corresponding di-
rect loading response is presented in Figure 5-26. Whereas the
support system capacity is roughly half (as expected), the remnant
capacity of the direct loaded bolting system is zero. Not only can a

9
As discussed in Section 5.3, a simple addition of each component’s full capacity would
only be valid if all components would reach their ultimate capacity at the same time.
However, in reality, when one component is at its full capacity, another component may
have failed or may only be at a small portion of its ultimate capacity.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


226 Rock support system capacity

system dissipate much less energy when directly loaded, the sys-
tems deformation capacity is drastically reduced.
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Figure 5-25 Indirect loading capacity of double-layer support system with 75 and 150
mm displacement limits for cables and threadbar, respectively; 200 mm central deflection
for surface support (= El Teniente mine assumption).

Figure 5-26 Direct loading capacity of double-layer support system with 37.5 and 75 mm
displacement limits for cables and threadbar, respectively; 100 mm central deflection for
surface support.
According to Tables 4-2 to 4-4, test data suggest that the direct
loading capacities are even lower than assumed for Figure 5-26.
The resulting characteristics are presented in Figure 5-27.
Following the approach introduced in this chapter, it is estimated
that the direct loading energy capacities of the bolting system and
the integrated support system are about 11 and 26 kJ/m2, respec-
tively. It would drop to zero at 75 mm even though the surface
support could theoretically have a remnant capacity of 17 kJ/m2.
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 227

This capacity cannot be mobilized because energy cannot be trans-


ferred to the bolts at this stage.
The indirect loading capacity of the bolting system is about 24
kJ/m2. It would drop to near zero (0.3 kJ/m2) at 75 mm and the
remnant surface support capacity of 17 kJ/m2 again cannot be mo-
bilized because loads and energy cannot be transferred to the bolts.

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
(a)

(b)
Figure 5-27 (a) Direct and (b) indirect loading capacity of double-layer support system
with recommended values from Table 4-2 with 20 and 15 mm displacement limits for
cables and threadbar, respectively; 100 mm central deflection for surface support.

The actual capacity, before the support system is pre-deformed


and severely damaged or becomes dysfunctional, is therefore in
the order of 15 to 24 kJ/m2 or about one quarter to one third of the
capacity estimated assuming simultaneous activation of all support
components. Because of energy splitting (between plate and rein-
forcement) the actual capacity will be higher and will fall some-
where between 24 and 40 kJ/m2. In Section 5.3.6, it was conclud-
ed that if the tendons were debonded (threadbar over 1.6 m and

Draft manuscript – Copyright protected – Cai and Kaiser 2018


228 Rock support system capacity

cables over 4 m) this support would at best dissipate between 31


and 45 kJ/m2.
Furthermore, after mining or seismic deformations have consumed
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

part of the displacement capacity, the remnant energy absorption


capacity will be substantially less; unless combined with debonded
cables, this double-layer support system has a relatively low ulti-
mate displacement capacity of about 75 mm.
Lessons learned from the El Teniente mine support system evolu-
tion over the last two decades are:
- the first layer with mesh-reinforced shotcrete is held by an initial
rockbolt pattern at 1 m ´ 1 m but the puncture capacity is likely
insufficient; and
- the second mesh layer over shotcrete is held by cablebolts with a
2 m ´ 2 m pattern through mesh and shotcrete.
The second layer reduces the spacing of the overall bolt patterns
and therefore increases the retention and reinforcement capacity of
the support system. The puncture capacity of the cables with
plates outside the double-layer system is much superior.
Because of the installation sequence, the second pass components
consume less capacity and this enhances the remnant capacity.
Most importantly, the double-layer system provides superior punc-
ture capacity and this ensures that the integrated support system is
able to dissipate energy over the entire displacement range.
Support consumption and sequence of support remediation
Even though discussions on site suggest that El Teniente mine has
not formally adopted a preventive support maintenance approach,
it is quite evident that they are implementing a proactive support
enhancement approach. For example:
- In the Dacita formation, known to be very brittle and thus highly
strainburst-prone, the two-layer support system with high
strength chain-link mesh is used throughout.
- In the extraction zone (high stress transition zone), the second
layer of the two-layer support system with high strength chain-
link mesh is only installed in areas of an identified need: fault
zones that experienced rockbursts in the past and in areas with
heavy cracking or delamination of shotcrete.
By implementing proactive support maintenance strategies, as dis-
cussed earlier in Section 5.3.2, today, mines can adopt state-of-
the-art support design methodology for support optimization to
ensure workplace safety.

5.4.6 Grasberg mine, Indonesia

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 229

Heavy strainbursting with event magnitudes reaching ML = 2 led


to repeated excavation damage at the drill and extraction horizon
during undercut advance at Grasberg mine in Indonesia. The gabi-
on concept was applied to design a burst-resistant support system.
This support performed well as it was able to retain the integrity of
the support system even though it often deformed beyond the
MPDL as illustrated by the images in Figure 5-19 and Figure 5-20
for Scenarios A to D.
Scenario D – failure of gabion by strainbursting behind the gabion
The integrity of the wall was maintained and unravelling between
tendons was mostly prevented (except in the lower portions of
some walls), however, the entire wall rotated or moved inward due
to heavy strainbursting inside and behind the gabion as illustrated
by the cartoons and photos in Figure 5-19.
In reality, the failure process is rather complex, as it can involve
rotational or translational failure modes driven by bulking (rock

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
volume increases), by tangential stress causing slip failure or by
energy releases of an unknown magnitude from the rock mass sur-
rounding the strainburst volume. If the failure is purely driven by
bulking, wall displacements will be limited to the depth of fractur-
ing involved and the associated bulking factors. For example, the
wall displacement due to bulking should be in the order of < 0.3 to
0.5 m (e.g., ≤ 100 mm due to ≤ 5% bulking inside the gabion plus
≤ 200 mm from 1 to 2 m of bulking behind the gabion at < 10%;
for a total of about 0.3 m). It is difficult to explain more than 0.4
m wall movement by bulking alone.
If caused by slip or shear rupture failure mechanisms, the kick-out
distance depends on the hanging wall (HW) to foot wall (FW) dis-
placement and wall displacements should be in the order of ≤ 0.2
to 0.5 m (i.e., for less than 1% HW/FW convergence).
If the failure is driven by energy release, the wall displacements
will depend on the energy consumed by the resistance forces of-
fered by the deep cablebolts and by the gabion. A simple model
was developed to assess the impact of adding more cables or add-
ing shear pins (see Volume II for details). This model suggests
that shear pins are most effective in increasing the gabion’s energy
dissipation capacity and that wall displacements can be reduced by
adding cablebolts, thereby increasing the energy dissipation capac-
ity of the support.
Scenarios A to C – failure of gabion by internal strainbursting
If bursting occurs inside the gabion, the following two scenarios
can lead to failure of the surface support or the reinforced rock
mass inside the gabion:
1) Direct loading scenarios (A and B) with burst within 0.5 to
1.5 m from wall; reduced capacity due to direct loading
Draft manuscript – Copyright protected – Cai and Kaiser 2018
230 Rock support system capacity

(this scenario is now mostly prevented, i.e., when effective


gabions are installed).
2) Indirect loading scenario (C) with burst within 1 to 2 m
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

from wall, i.e., inside the reinforced gabion.


The following simplifications are made to conduct a generic ca-
pacity assessment, following the support system capacity model
introduced in Sections 5.3.3 and 5.3.4.
i. Retention system:
- Standard mesh and bolting; or
- Mesh-reinforced shotcrete; or
- Double-layer scenario for stable gabion.
ii. Baseline bolting system:
- 2.1 m long 22 mm rebar (or D-bolts) at 1 m ´ 1 m plus
74 mm fibre shotcrete (ignored) and 8 mm diameter
mesh installed at face (0 mm wall displacement); 1 re-
bar/m2;
- 6 m long 15.2 mm single plane cables at 2 m ´ 2 m
Ring 1 installed at 10 mm wall displacement; and
- 6 m long 15.2 mm single plane cables at 1 m ´ 2 m
Ring 2 also installed at 10 mm wall displacement.
iii. Secondary support system installed at 30 or 75 mm wall
displacement
- add 15.2 mm plane cable at 2 m ´ 2 m; and
- add 22 mm D-bolt or 32 mm threadbar at 0.5 bolts/m2;
or
- add double cable at 1.14 m spacing.

Direct loading scenario (A) by strainburst near wall


In this situation, the retention system and the bolts are directly
loaded at the plate and the model shows that the baseline support
can survive direct loading from strainbursts up to 12.7 kJ/m2 as
long as the bolt displacement is less than 30 mm. However, the
gabion will unravel if mesh is used.
The capacity of the baseline support is doubled when mesh-
reinforced shotcrete is used as long as the displacements are less
than 50 and 95 mm for threadbar and D-bolts, respectively. Theo-
retically, this enhanced gabion with supplemental bolts and mesh-
reinforced shotcrete unravels when the strainburst energy exceeds
32.1 kJ/m2, although very high bolt forces will likely cause prema-
ture, puncture failure of the retention system.

Indirect loading scenario (B) by strainburst inside the reinforced rock


mass
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 231

If the retention system is stable during direct loading, i.e., it does


not fail during a strainburst, the bolts are loaded indirectly thereby
enhancing the system’s energy and displacement capacities.
In this case the model shows that the capacity of the baseline sup-
port is quadrupled as long as the displacement is less than 180 mm.
The bolt forces are still excessive but the reinforced rock will
move with the surface support and provide a much-improved sur-
face support/rock mass package. This system demonstrates a good
remnant capacity.
This example explains how the gabion works and why the gabion
concept provides a superior rockburst support system.

Role of PSM for indirect loading scenario


Before PSM the baseline support with threadbar (but without D-
bolts) has an energy capacity of 36.6 kJ/m2 but loses most of it at
75 mm of deformation.
If PSM is applied to this support using double cables installed at

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
100 mm of wall displacement, the remnant energy capacity has
increased substantially (from 9.7 to 26.5 kJ/m2 after 70 mm of
wall displacement).
It must be noted that the effectiveness of PSM is sensitive to the
timing of support remediation. Quite often it is impossible to con-
duct a PSM at an optimal stage. For a scenario with a delayed in-
stallation of double cables (after 175 mm of wall displacement)
through a stable gabion wall, it can be shown that a substantial
remnant energy capacity (17.8 kJ/m2) is maintained between 125
and 225 mm of wall movement. However, the load capacity is se-
riously compromised between 150 and 175 mm of wall displace-
ment. The intact gabion, nonetheless, provides a stable workplace
as long as no further strainbursting occurs.

Apparent indirect loading capacity


If all support components, including D-bolts, were installed be-
tween 0 to 50 mm of deformation, the support system would have
an apparently great characteristic with an allowable displacement
limit of 180 mm, a maximum energy capacity of 71.6 kJ/m2 and a
remnant capacity after 90 mm or almost 40% (28.2 kJ/m2). How-
ever, at 90 mm of deformation the energy capacity of the bolting
system would drop to about 25% of the maximum and the load
capacity to about 25% of the maximum at 105 mm of deformation.
In other words, the full capacity is only available if the support
system has not been deformed before a strainburst occurs. If the
support has been deformed by about 100 mm, the actual (remnant)
capacity is much less, about 18 kJ/m2 instead of 72 kJ/m2.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


232 Rock support system capacity

These examples illustrate how each support component contrib-


utes to the load and energy capacities of a support system and how
the cumulative capacity can be consumed by mining-induced de-
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

formation.

5.4.7 Comments on other case examples presented in the


CRBSHB
Several case histories are presented to illustrate the nature of rock-
burst damage in the CRBSHB’96. These case histories have to be
interpreted in the light of the new concepts introduced in this book.
It is quite evident that many of the damage cases involve a signifi-
cant or dominant strainburst component. Comments and interpre-
tations presented in the CRBSHB’96 must therefore be carefully
considered and revised.

5.5 References

Andrieux, P., Turichshev, A., O’Connor, P., and Brummer, R.K. 2005.
Dynamic testing with explosive charges of rockburst-resistant ground
support systems at the Fraser Nickel Mine. Report to Falconbridge
Limited Mine Technical Services. p.
Brändle, R., Rorem, E., Luis, R., and Fisher, R. 2017. Full-scale dynamic
tests of a ground support system using high-tensile strength chain-link
mesh in El Teniente mine, Chile. In 1st Internation Conf. on
Underground Mining Technology ACG. pp. 25-43.
Bucher, R., Cala, M., Zimmerman, A., Balg, C., and Roth, A. 2013. Large
scale field tests of high-tensile steel wire mesh in combination with
dynamic rockbolts subjected to rockburst loading. In Ground Support.
Edited by B.G.H. Brady and Y. Potvin. pp. 221-232.
Counter, D. 2010. Geotechnical Mine Design Package and 2009 Annual
Mine Stability Review, Xstrata Copper Kidd Creek Mine. p.
Diederichs, M.S. 2014. When does Brittle Failure Become Violent? Spalling
and Rockburst Characterization for Deep Tunneling Projects. In
Proceedings of the World Tunnel Congress. pp. 1-10.
Espley, S. 1999. Thin spray-on liner support and implementation in the
hardrock mining industry. Laurentian University. Master Thesis. p. 311.
Falmagne, V., Anderson, T., Conlon, B., and Judge, K. 2005. Impact testing
of prototype MCB33 and MCB38 bolts, Unpublished Canment MMSL
report (05-007) (CR), Canada. p.
Gaudreau, D., Aubertin, M., and Simon, R. 2004. Performance assessment
of tendon support systems submitted to dynamic loading. In Proc 5th Int
Symp on Ground Support. Edited by E. Villaescusa and Y. Potvin.
Balkema, Perth, Australia. pp. 299-312.
Hadjigeorgiou, J., and Potvin, Y. 2007. Overview of dynamic testing of
ground support. In Deep Mining 07. Edited by Y. Potvin. Australian
Centre for Geomechanics, Perth. pp. 349-371.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 233
Hagan, T.O., Milev, A.M., Spottiswoode, S.M., Hildyard, M.W., Grodner,
M., Rorke, A.J., Finnie, G.J., Reddy, N., Haile, A.T., Le Bron, K.B., and
Grave, D.M. 2001. Simulated rockburst experiment - an overview. J.
South Afr. Inst. Min. Metall.: 217-222.
Haile, A.T., and Le Bron, K. 2001. Simulated rockburst experiment
evaluation of rock bolt reinforcement performance. J. South Afr. Inst.
Min. Metall.(August): 247-251.
Heal, D., Hudyma, M., and Potvin, Y. 2004. Assessing the in-situ
performance of ground support systems subjected to dynamic loading.
In Ground Support in Mining and Underground Construction. pp. 319-
326.
Hedley, D.G.F. 1992. Rockburst handbook for Ontario hardrock mines.
CANMET Special Report SP92-1E. p. 305.
Hildyard, M.W., and Milev, A.M. 2001. Simulated rockburst experiment:
Development of a numerical model for seismic wave propagation from
the blast, and forward analysis. J. South Afr. Inst. Min. Metall.: 235-245.
Hoek, E., and Brown, E.T. 1980. Underground excavations in rock.
Institution of Mining and Metallurgy, London. p. 527.

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Hoek, E., Kaiser, P.K., and Bawden, W.F. 1995. Support of Underground
Excavations in Hard Rock. A.A. Balkema. p. 215.
Hutchinson, D.J., and Diederichs, M.S. 1996. Cablebolting in Underground
Mines. BiTech Publishers Ltd. p. 416.
Jager, A.J. 1992. Two new support units for the control of rockburst damage.
In Rock Support in Mining and Underground Construction. Edited by
P.K. Kaiser and D.R. McCreath. pp. 621-631.
Joo, A.M. 2017. Dynamic ground support at El Teniente Mine. Personal
communication with P.K. Kaiser. RASIM9 short course presentation at
RASIM9.
Kaiser, P.K. 2017. Excavation vulnerability and selection of effective rock
support to mitigate rockburst damage. In Rockburst: Mechanism,
Monitoring, Warning and Mitigation. Edited by X.-T. Feng. Elsevier. pp.
473–518.
Kaiser, P.K., Tannant, D.D., and McCreath, D.R. 1996. Canadian Rockburst
Support Handbook. Geomechanics Research Centre, Laurentian
University, Sudbury, Ontario. p. 314.
Kirsten, H.A.D., and Labrum, P.R. 1990. The equivalence of fibre and mesh
reinforcement in the shotcrete used in tunnel-support systems. J. South
Afr. Inst. Min. Metall. 90(7): 153-171.
Li, C. 2010. A new energy-absorbing bolt for rock support in high stress
rock masses. Int. J. Rock Mech. Min. Sci. 47: 396-404.
Ortlepp, W.D. 1969. An empirical determination of the effectiveness of
rockbolt support under impulse loading. In Proc, Int. Symp. on Large
Permanent Underground Openings. Edited by B. Jorstad, Oslo. pp. 197-
205.
Ortlepp, W.D. 1992. The design of support for the containment of rockburst
damage in tunnels an engineering approach. In Rock Support in Mining
and Underground Construction. Edited by P.K. Kaiser and D.R.
McCreath. Balkema. pp. 593-609.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


234 Rock support system capacity
Ortlepp, W.D., and Erasmus, P.N. 2002. The performance of yielding straps
under dynamic loading. In 2nd Int. Seminar on Surface Support Liners.
SAIMM, Sandton.
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Ortlepp, W.D., and Swart, A.H. 2002. Performance of various types of


containment support under quasi-static and dynamic loading conditions,
Part II. p. 100.
Player, J.R. 2004. Field performance of cone bolts at Big Bell mine. In
Ground Support in Mining and Underground Construction. Edited by E.
Villasescusa and Y. Potvin. pp. 289-298.
Player, J.R., Morton, E.C., Thompson, A.G., and Villasescusa, E. 2008.
Static and dynamic testing of steel wire mesh for mining applications of
rock surface support. In 6th International Symposium on Ground Support
in Mining and Civil Engineering Construction. pp. 693-706.
Potvin, Y., and Heal, D. 2010. Dynamic testing of High Energy Absorption
(HEA) mesh. In 5th International Seminar on Deep and High Stress
Mining. Edited by M. Van Sint Jan and Y. Potvin, Santiago, Chile. pp.
283-300.
Potvin, Y., Wesseloo, J., and Heal, D. 2010. An interpretation of ground
support capacity submitted to dynamic loading. In 5th International
Seminar on Deep and High Stress Mining. Edited by M. Van Sint Jan
and Y. Potvin, Santiago, Chile. pp. 251-272.
Shirzadegan, S., Nordlund, E., and Zhang, P. 2016. Large scale dynamic
testing of rock support system at kiirunavaara underground mine. Rock
Mechanics and Rock Engineering 49(7): 2773-2794.
Simser, B., Andrieus, P., and Gaudreau, D. 2002a. Rockburst support at
Noranda’s Brunswick Mine, Bathurst, New Brunswick. In NARMS
2002. Edited by R. Hammah and W. Bawden and J. Curran and M.
Telesnicki. University of Toronto Press. 1 pp. 805-813.
Simser, B., Joughin, W.C., and Ortlepp, W.D. 2002b. The performance of
Brunswick Mine’s rockburst support system during a severe seismic
episode. J. South Afr. Inst. Min. Metall.: 217-223.
Stacey, T.R., and Ortlepp, W.D. 2001. Tunnel surface support capacities of
various types of wire mesh and shotcrete under dynamic loading. J.
South Afr. Inst. Min. Metall.: 337-342.
Tannant, D.D., Kaiser, P.K., and McCreath, D.R. 1995. Large-scale impact
tests on shotcrete. p. 45.
Turner, M.H., and Player, J.R. 2000. Seismicity at Big Bell Mine. In
Proceedings Massmin 2000. The Australasian Institute of Mining and
Metallurgy, Melbourne. pp. 791-797.
Villaescusa, E., de Zoysa, U., Player, J.R., and Thompson, A.G. 2016.
Dynamic testing of combined rockbolt and mesh schemes. In Procc.
Seventh International Conference & Exhibition on Mass Mining. The
AusIMM, Sydney, Australia. pp. 1-10.
Yao, M., Chinnasane, D.R., and Harding, D. 2009. Mitigation plans for
mining in highly burst-prone ground conditions at Vale Inco Copper
Cliff North Mine. In ROCKENG09: Proceedings of the 3rd CANADA-
US Rock Mechanics Symposium. Edited by M. Diederichs and G.
Grasselli, Toronto. Paper 4045.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 235

Appendix A: Terminology
Excavation damage and rockburst types
Rockburst A rockburst is defined as damage to an excavation that occurs in a
sudden and violent manner and is associated with a mining-induced
seismic event. “Rockburst” is a generic term and is independent of
the cause of damage and failure process. Strain-, pillar- and fault
slip-bursts are all rockbursts if they cause damage to an excavation
or its support. The seismic event may be remote from or co-located
with the damage location. A seismic event alone without causing
damage is not a rockburst.
Excavation Excavation damage is visible damage to an excavation in the form

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
or rockburst of overbreak that is caused by a dynamic event if it is vulnerable
damage (see below). It is reflected in many forms of damage such as stress
fracturing causing rock mass unraveling and falls of ground or
damage to the rock support in the form of support damage (crack-
ing, bagging, bolt failure). Such damage can be caused by excessive
stress, by shaking marginally stable ground or by energy transferred
from remote seismic (fault slip) events. This definition of damage
differs from the terminology adopted in the nuclear waste manage-
ment literature where rock damage in an excavation damage zone
(EDZ) includes crack damage and damage that does not lead to
excavation instability.
Damage with rock or support component ejection may not or only in
part be related to the intensity (magnitude, moment or energy) of a
remote or co-located seismic event.
Excavation Excavations are more or less vulnerable to damage depending on
vulnerability how close the stress near an excavation is to the point of failure (i.e.,
to the local rock mass strength. In general, the higher the static
factor of safety or the lower the static probability of failure, the
lower the vulnerability of an excavation to rockbursting. Conse-
quently, the excavation vulnerability depends on the rock failure
type and it increases with increasing stress level (stress/strength),
proximity to geological structures, the local excavation deformation
potential (reflected in the local mine stiffness, extraction ratio, and

Draft manuscript – Copyright protected – Cai and Kaiser 2018


236 Appendix A Terminology

excavation geometry; see below), and the effectiveness of the sup-


port system (its ability to dissipate energy in the reinforced rock
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

arch). Note: support components such as shotcrete may become


vulnerable to failure even if an excavation is stable, e.g., due to
stress wave reflections.
Excavation When part of an excavation fails causing overbreak, the entire exca-
deformation vation deforms in response to the related change in excavation
potential geometry. The associated deformation pattern depends on the sur-
rounding rock mass properties, the local stress condition, the exca-
vation geometry, and the shape of the change in geometry. Condi-
tions that lead to larger deformations during a change in geometry
(caused by rockburst damage) have an elevated excavation defor-
mation potential. For example, a notch in a tunnel wall at a T-
intersection will have a higher deformation potential than an identi-
cal notch away from the intersection. A dyke imbedded in a soft
rock mass will have a higher deformation potential than that imbed-
ded in a stiff rock mass. There are many other factors such as the
local mine stiffness (see next) that affect the deformation potential.
Loading system The local mine stiffness describes the response of a rock mass to
stiffness LSS rock failure. The LSS is the ratio of an induced stress change and
or local mine the related strain increment in a rock volume (or the ratio of an
stiffness induced force change and the related displacement, e.g., during a
pillar failure). Similar to a laboratory test frame, the loading system
in a mine opening can be soft or stiff. In a stiff environment, a given
strain increment will cause a large stress change (drop) and in a soft
environment, the same strain increment will cause little stress
change. Hence, stiff loading systems impose less energy on failing
rock than soft systems.
Tarasov and Potvin (2013) refer to the effect of LSS as a cause for
“relative brittleness”, i.e., brittle failure resulting from external
system factors. They use the term “intrinsic brittleness” in the con-
text of laboratory tests for brittleness that is solely related to the
rock properties.
Strainburst A strainburst is a sudden and violent failure of rock near an excava-
tion boundary caused by excessive straining of a volume of stiff and
strong rock (burst volume). The primary or a secondary seismic

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 237

source is co-located at the damage location.


A self-initiated strainburst is a rockburst caused by a gradual weak-
ening of the rock mass such that the local stress after some time
reaches the rock mass strength in a relatively soft loading/mining
system, i.e., the local mine stiffness has to be softer than the post-
peak behavior of the failing rock mass. In this case, the damage is
only related to the energy stored in the failing rock volume and the
energy released from the surrounding rock mass. The radiated ener-
gy or intensity of the seismic event is only related to the strainburst
intensity.
A mining-induced strainburst is a rockburst caused by mining-
induced deformations or strain (due to tunnel advance, stope exca-
vation, cave propagation or loading, etc.) that change the local stress

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
near an excavation such that the stress (temporarily) reaches the
rock’s strength and the actual rock strain exceeds the strain at the
peak strength. The radiated energy or intensity of the seismic event
is again only related to the strainburst intensity.
A seismically triggered strainburst is a self-initiated or a mining-
induced strainburst that is triggered by a remote seismic event. In
this case, the remote seismic event is the primary seismic event and
the seismic event co-located with the strainburst damage is a sec-
ondary event. However, the damage is not related to the intensity of
the remote seismic event, it only serves as the trigger of the strain-
burst.
A dynamically loaded strainburst is a strainburst that is augmented
by the impact of energy radiated from a primary source in two pos-
sible forms:
- the radiating energy causes a dynamic stress pulse that may
deepen the depth of failure, thus releasing more stored energy,
and through rock mass bulking adding additional strain or dis-
placement to the rock and support; or
- the radiated energy may transfer some of its radiated energy to
kinetic energy and eject part of marginally stable rock.
A strainburst is only associated with rock ejection if some of the
released energy is transferred in the form of kinetic energy to non-

Draft manuscript – Copyright protected – Cai and Kaiser 2018


238 Appendix A Terminology

supported or poorly supported rock.


A strainburst may also occur in a well-supported rock mass, i.e.,
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

behind mesh or shotcrete, in the reinforced rock or behind the sup-


ported ground. This is called a restrained self-initiated, mining-
induced, triggered or dynamically loaded strainburst. Such re-
strained strainburst may cause damage to the support system and
cause shotcrete ejection.
Pillar burst A pillar burst is a sudden and violent failure of rocks in the pillar
core or the complete collapse of a pillar, whether in a room, post,
crown, rib, or sill pillar. In a pillar burst, the source and damage are
co-located, i.e., damage is located in the confined pillar core. A
pillar burst may be triggered by a remote seismic event but the
damage is dominated by the energy released from the failing pillar.
Fault-slip burst A sudden and violent failure or damage to an excavation caused by
the dynamic slippage along a pre-existing fault or along a newly
generated shear rupture is called a fault-slip burst. The damage to an
excavation is then strongly related to the energy radiated from the
seismic event at the slip or rupture location but some energy will
also be released from the failing rock or concrete near the excava-
tion.

Rock mass damage mechanisms


Bulking of stress- Bulking is a term borrowed from blasting or fragmentation. It
fractured ground describes an increase in volume due to the transition from a compe-
tent rock mass to a volume of broken rocks. The bulking of rock is
largely a result of geometric non-fit of very strong rock fragments
caused by stress fracturing under static or dynamic loading. In a
rockburst, this bulking process occurs in a sudden and violent
manner.
Fall of ground or A fall of ground (FoG) is a gravity-driven damage process whereby
shakedown a volume of marginally stable ground fails. The failing volume of
rock may be defined by geological structures (discontinuities)
forming single or groups of rock wedges or a zone of fractured
rocks.
A FoG may be seismically triggered or dynamically loaded. In the

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 239

latter case, vibrations or shaking from a remote seismic source


increases the demand on the support by adding dynamic accelera-
tion to the gravitational acceleration. A FoG triggered or dynami-
cally loaded by a dynamic disturbance is thus called a shakedown.
Rock ejection Rock ejection is a dynamic process driven by some source of kinet-
ic (not potential) energy. This kinetic energy may stem from a
bulking process, from momentum transfer from larger to smaller
rocks, or energy transfer by seismic radiation of energy from a
remote seismic event.
Rock ejection, therefore, is related to the remnant kinetic energy,
after all other energy sinks have extracted energy that is available
for transfer to rock blocks.
In many situations, the ejection velocity is not directly related to

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
ground motion intensity, or the PGV (peak ground velocity). The
common assumption of setting the ejection velocity equal to the
PGV is therefore rarely valid for burst-resistant support design. The
ejection velocity is often in part or entirely related to the locally
releasable strain energy and the kinetics of the failing rocks and
concrete.
Qualifiers
Induced or driven The qualifiers induced or driven relate to a source of a problem
without specifying what the actual mechanism and effect is. For
example:
Mining-induced means that without mining nothing would happen.
Stress-driven means that failure is caused by high stress or stress
relaxation.
Gravity-driven indicates that gravity or potential energy provides
the main source of energy.
Deformation-driven indicates that imposed strains or displacements
provide the main source of loading.
Energy-driven indicates that released and radiated energies provide
the main source of energy to cause failure.
Triggered This qualifier defines that some disturbances only initiate or trigger

Draft manuscript – Copyright protected – Cai and Kaiser 2018


240 Appendix A Terminology

failure but the dynamic disturbance does not add damage-causing


stress, deformation or energy.
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Loaded or This qualifier defines a process whereby load or stress is added (or
stressed removed/relaxed) and more critical conditions are created to pro-
mote and propagate failure; i.e., substantial stress, deformation or
energy is added.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 241

Appendix B: Nomenclature
𝑎𝑎⃗ Acceleration
a Radius of a circular tunnel
*
a Empirical constant in the scaling law
ASTM American Society for Testing and Materials
BF Bulking factor = change in radial length / length; the average
BF over the depth of failure is the change in radial length / df
BurstSupport A tool for assessing rock support demand based on the
methodology of the Canadian Rockburst Support Handbook
(Kaiser et al. 1996)
C Capacity in terms of load, displacement, or energy; subscripts
“s” for static, “d” for dynamic, and “R” for remnant capacity

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
C* Empirical constant in the scaling law
CANMET Canada Centre for Mineral and Energy Technology
CFS Central factor of safety
CoV Coefficient of Variation = standard deviation / mean
cs S-wave velocity
cp P-wave velocity
CRBSHB … Canadian Rockburst Support Handbook (Kaiser et al. 1996)
D Demand in terms of load, displacement, or energy; subscripts
“s” for static and “d” for dynamic
DC Double couple source model
DEM Distinct Element Method
DFF Depth of failure factor (= E3/E4)
DIF Damage initiation factor (= E1/E2)
df Depth of failure in unsupported ground; superscripts “m” for
mean (dfm) and “e” for extreme (dfe)
dfo Depth of failure before a strainburst occurs behind fractured
rock
dSB Depth of strainburst behind a statically fractured zone
DP Deformation potential
Es Radiated seismic energy

Draft manuscript – copyright protected – Cai and Kaiser 2018


242 Nomenclature

E Energy related rock support; subscripts “p” for peak energy,


“ult” for ultimate energy, 100 and 200 for energy at displace-
ment thresholds of 100 and 200 mm, “cs” for consumed energy,
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

“system” for rock support system energy, “bolt plate” for rock-
bolt energy through loading at plate, “reinforcement” for rock-
bolt energy through split loading, “surface” for surface support
element energy, and “rock mass” for energy consumed by frac-
turing the rock
Erm Rock mass modulus
Es Radiated seismic energy
EVP Excavation vulnerability potential index
F Force
f Frequency
f0 Corner frequency
FEM Finite Element Method
FoG Fall(s) of ground
FS Factor of safety; subscripts “s” for static, “d” for dynamic,
“Load” for force, “Displ” for displacement, and “Energy” for
energy equilibrium
g Gravitational acceleration
GRC Ground reaction curve (also: Geomechanics Research Centre at
Laurentian University of Canada)
GSI Geological strength index
l Wave length
le Yield length of rockbolt
L Load: subscripts “p” for peak load, “ult” for ultimate load, and
100 and 200 for average load to displacement thresholds of 100
and 200 mm
Lg Gradual deformation zone length of rockbolt
Leff Effective length of a beam
Lfs Free stretching length of rockbolt
LSS Loading system stiffness in [force/deformation] or
[stress/strain]; LSS* for critical LSS
k Stress ratio (horizontal to vertical)
m Mass
M0 Seismic moment

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 243

MCB Modified conebolt


MEL Maximum energy limit
Mi,j (i, j = 1, 2, 3) Failure modes in the excavation behaviour
matrix
ML Local or Richter magnitude
MN Nuttli magnitude
MPDL Maximum practical displacement limit
MPEL Maximum practical energy limit, previously called MPSL
MPSL Maximum practical support limit
Pf or u Probability of failure or undesired behavior
PGA Peak ground acceleration
PGD Peak ground displacement
PGV Peak ground velocity. Subscripts “D” for design PGV values

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
estimated from the scaling law, “s” for S-wave PGV
PPSfrm Post-peak strength of reinforced, fractured rock mass
PSM Proactive or preventive support maintenance
Q Barton’s rock mass quality index; Q’ = Q with Jw/SRF = 1
R Distance from seismic source in scaling law
R1 to R5 Rockburst damage scale
RBP Rockburst potential
RBS Rockburst severity
RMQ Rock mass quality (1 to 3) used in excavation behaviour matrix
RMR Bieniawski’s rock mass rating index
r Radius of gyration
s Bolt spacing
SCC Support capacity consumption
S-GMAT Synthetic Ground Motion Assessment Tool from Institute for
Mine Seismicity (IMS)
SBP Strainburst potential
SBS Strainburst severity
SL Stress Level index; SL = smax/ UCS with smax = 3s1 – s3
SM Safety margin; subscripts “s” for static and “d” for dynamic;
SM = 1 – (1/FS)

Draft manuscript – Copyright protected – Cai and Kaiser 2018


244 Nomenclature

SRF Strength reduction factor (in the Q-system)


SPECFEM Spectral Finite Element Method
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

t Thickness
Dt Rupture time
UCS Uniaxial compressive strength
v Phase speed of wave
V Volume; subscript “SB” for strainburst volume
vej Ejection velocity when rock fragments detach from the rock
mass or concrete detaches from the support system
vi Initial ground velocity = highest velocity of damaged, frag-
mented rock mass (vi ≥ vej; vi ¹ PGV)
W Work or energy; subscripts “r” for released, “f” for fracturing,
“fs” for fracturing of supported rock, “ys” for yielding support,
“h” for heat, “ej” for ejected fragments, and “k” for kinetic (Wk)
WASM Western Australian School of Mines
x Return period of loading
z Overburden depth
a1, a2 Energy sharing coefficients for rockbolt and surface support
d Deformation; subscripts “B” for bolt deformation and “S” for
surface support component deformatin
dall Allowable deformation limit from an operational perspective
dall Maximum deformation capacity of a rock support component
dult Ultimate deformation limit from a design extremity perspective
dp Displacement of rockbolt corresponding to its peak load
l Spring stiffness
s1, s3 Major and minor principal far-field stresses
smax Maximum excavation-induced stress on the wall of circular
tunnel in elastic rock; smax = 3 s1 – s3
stens Tensile strength
r Density

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 245

Appendix C: Static and dynamic


capacities of rockbolts
A summary of published rockbolt support capacities is presented in the following table. It
contains data from currently available bolt types independent of their practical use and
independent of data quality controls. Missing data are left blank. This table does not provide
recommended design values and the user is encouraged, as discussed throughout this book, to
critically evaluate the applicability of the underlying test before adopting any of the listed values
for design purposes.
Note: Lp = peak load of bolt, Lult = load at ultimate displacement, Eu = ultimate energy capacity of
bolt, dult = ultimate displacement capacity of bolt.

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Draft manuscript – Copyright protected – Cai and Kaiser 2018


246 Appendix C

Table C-1 Static and dynamic capacities of rockbolts (for definition of terms refer to Figure 4-1d)
Static Dynamic
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Bolt type Lp dult Eu Lult dult Eu Note


(kN) (mm) (kJ) (kN) (mm) (kJ)

Cablebolt 12.7 mm un-de- 150– Fully grouted to collar


183 20–40 2–6 20–40 2–6 a
bonded 180 a – inferred from static capacity

Cablebolt 15.2 mm un-de-


260 20–40 3–6 300 26 7.8
bonded

a – dynamic device yielding

Cablebolt 15.2 mm 150– 80– b – 300 mm (standard); larger capacity may


300 > 30 300 b 30 c
Dynamic cablebolt 180 120a be achievable

c – at 300 mm displacement capacity

4 m de-bonding length
Cablebolt 15.2 mm 200– a b
140 28 200 140 25–30 a – based on min. 3.5% elongation limit
4 m de-bonded 250
b – inferred from static capacity

Cablebolt 15.2 mm 70– a – Maximum load 100 kN, steady-state 70


> 200 > 14 > 220 > 19
Duracable 100 a kN

Conebolt 16 mm South a – inferred from quarry block blast test


100 > 200 > 20 > 20 a
Africa (S.A.) result

a – inferred from quarry block blast test


Conebolt 22 mm S.A. 200 > 200 > 40 39.2 a
result

a – 355 mm free stretching length;

200– b – double embedment split dynamic test


D-bolt 20 mm 160 50–60 a 8 50–53 13.1 b
230 method; bolt did not fail in one impact

a – per meter of free stretching length


D-bolt 22 mm 200 50–60 250 150/m a 36/m b b – double embedment split dynamic test
method;

8 per
Durabar 16 mm 100 > 600 48 100 ~ 600 100 mm a – per 100 mm sliding
a

Friction set 39 mm 50–80 80–250 5–15 2.4 m long

5 per 2.4 m long


2.7±1
70– 100 25-
Friction set 46 mm 80–250 300–600 per 100 a – per 100 mm sliding in one meter
100 mm/m 30/m
a mm/m embedment

a – locked in load capacity (sliding capacity


100–
Garford solid bar 21.7 mm 215 a ~500 ~ 500 b 27–33 c is 140 kN)
125
b – total displacement capacity up to 500

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 247
Static Dynamic
Bolt type Lp dult Eu Lult dult Eu Note
(kN) (mm) (kJ) (kN) (mm) (kJ)

mm may be specified

c – direct impact to plate

MCB33 conebolt 17.3 mm 100–


110 150 20 63–251 > 16 Grease debonded
greased 170

100–
MCB38 conebolt 17.3 mm 110 150 15–20 65–250 > 16 Grease debonded
200

MCB33 conebolt 17.3 mm 100– 100– PVC debonded


150–200 15–25 300–600 30–35a
PVC debonded 110 150 a – direct impact to plate

Mechanical bolt 17.3 mm 143 20–50 2–6 11–40 1–5 C1055 MOD steel

Rebar 20 mm, resin Grade 400 W steel


116 10–20 a 1–4
grouted a – pulled/failed at thread

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Grade 400 W steel
2.2–4.0
Rebar 22 mm, resin a a – direct impact to plate
160 10–30 2–5 1–10
grouted
12–14 b b – double embedment split with a 50 cm
stretch length

Rebar 25 mm, resin


210 10–30 3–7 Grade 400 W steel
grouted

5.4 per a – direct impact to plate, per 100 mm plate


20 per displacement
Roofex (Rx20D) 200 150–800 100 100 mm
100mm
+ 7.5 a

a – direct impact to plate, per 100 mm plate


3.0 per displacement
8 per 25b
Roofex (Rx8D) 80 150–800 55 100mm
100mm mm/kJ b - The displacement rate is very much a
+6.7a
constant

a – inferred from quarry block blast test


Smooth bar 16 mm fully result (Ortlepp 1994)
120 50–100 5–10 380 >16 a, b
grouted b – (3.0L-0.15) kJ/100 mm sliding, where L
is the initial embedded length in meter

2.2–6.6 S355 JR steel


per
Swellex, Mn 12 yielding 100 31-113 a – double embedment split dynamic test
100mm/
ma method. Tube split is 500mm from the plate.

6.2±1 a – per m embedment length


a per
Swellex, Mn 24 yielding 70/m 50–60 50/m 300–600 b – double embedment split dynamic test
100mm/
mb method

Threadbar 20 mm 146 20–70 3–11 155– 8–15 0.8–1.1 Grad75 steel

Draft manuscript – Copyright protected – Cai and Kaiser 2018


248 Appendix C
Static Dynamic
Bolt type Lp dult Eu Lult dult Eu Note
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

(kN) (mm) (kJ) (kN) (mm) (kJ)


a
196 a – direct impact to plate; likely failed at
thread

Threadbar 20 mm 13.6– a – double embedment split dynamic test


146 20–70 ≤11 50–100
0.16 m debonded 21.8 a method (160 cm debonding length)

a – bolt did not fail. The energy absorption


110– rate is 16.4 kJ per 200 mm of displacement
Yield-Loc 17.2 mm 110 > 200 > 20 > 200 16.4 a, b
160
b – direct impact to plate

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 249

Appendix D: Static and dynamic capacities


of surface support components
The following table contains a summary of published test data for retention systems. It contains data
from all currently available surface support types independent of their practical use and independent of
data quality controls. Missing data are left blank. This table does not provide recommended design
values and the user is again encouraged to critically evaluate the applicability of the underlying test
before adopting any of the listed data.
Note: Lp = peak load of bolt, Lult = load at ultimate displacement, Eu = ultimate energy capacity of mesh
or shotcrete in test (it may not represent the ultimate energy capacity), dult = central displacement of
mesh or shotcrete at end of test (it may not represent the ultimate displacement capacity).

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Draft manuscript – Copyright protected – Cai and Kaiser 2018


250 Appendix D

Table D-1 Summary of published test data for mesh support elements (summarized from Kaiser et al. (1996) and Stacey and
Ortlepp (2001))
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Static Dynamic
Mesh type Lp dult Eu Lult dult Eu Note
(kN) (mm) (kJ/m2) (kN) (mm) (kJ/m2)

#9-gauge weld mesh GRC test facility, 3.658 mm


125–
12–18 1–4 wire diameter; 1.2 m x 1.2 m
175
square pattern

#6-gauge weld mesh 150–


24–28 4–6 10 GRC, 4.877 mm
225

#4-gauge weld mesh 175–


34–42 6–9 55 200 GRC, 5.893 mm
250

#9-gauge chain-link mesh > 400–


32–38 10–12 15
450

Special mesh 300– South African (S.A.) test


30–35
400 facility, high strength steel

TECCO® chain-link mesh a – simulated using a


150 185 300 50 a
(Geobrugg mesh) (4 mm) numerical tool

With cable lacing

HEA mesh (weld mesh with No dynamic test results


pre-fabricated cable lacing)
There is a high potential for
150– 800– the integrated lacing to cause
30–40 high, unpredictable shear and
170 900
bending loads at bolt heads. It
is not recommended by the
authors; see Appendix E.

Chain-link mesh + rope S.A.


lacing
b – due to input energy loss in
300 35 b deforming the bricks, these
capacity values may need to
be reduced for design

Weld mesh + rope lacing 300–


40 S.A.
400

Special mesh + rope lacing 450 50 S.A.

With mesh strap Distributed load tests

#6-gauge weld mesh + #0- 100– S.A., behaviour will depend


40–50
gauge mesh straps 200 on pattern of straps

Chain-link mesh + #0- 100– 40–50 S.A., behaviour will depend

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 251

Static Dynamic
Mesh type Lp dult Eu Lult dult Eu Note
(kN) (mm) (kJ/m2) (kN) (mm) (kJ/m2)

gauge mesh straps 200 on pattern of straps

#4-gauge weld mesh + #0- S.A., behaviour will depend


gauge mesh straps on pattern of straps. First
100–
> 50 impact. The energy is
200
absorbed by the system, not
just the surface support alone.

High tensile strength mesh S.A., behaviour will depend


+ Durastraps on pattern of straps. First
100–
> 50 impact. The energy is
200
absorbed by the system, not
just the surface support alone.

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Draft manuscript – Copyright protected – Cai and Kaiser 2018


252 Appendix D

Table D-2 Summary of published test data for shotcrete support elements (summarized from Kaiser et al. (1996), Stacey and
Ortlepp (2001), and Potvin et al. (2010))
Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Static Dynamic

Shotcrete type Lp dult Eu Lult dult Eu Note


2 2
(kN) (mm) (kJ/m ) (kN) (mm) (kJ/m )

Un-reinforced shotcrete < 50 << 5 GRC test facility

Fiber-reinforced shotcrete
20–30 1–3 <5 GRC
(50 to 70 mm thickness)

Steel fiber-reinforced
shotcrete (50 to 70 mm 20–30 NA 1–3 30–40 5 GRC
thickness)

Dramix fiber-reinforced South African (S.A.)


100 ≤ 10
shotcrete distributed load test facility

Monofilament
South African (S.A.)
polypropylene fiber- 120 ≤ 15
distributed load test facility
reinforced shotcrete

#6-gauge gauge weld mesh- 100–


45–55 6–9 150 15 GRC
reinforced shotcrete 150

With cable lacing

Fiber-reinforced shotcrete 200 to ≤ 15 to South African (S.A.)


+ rope lacing 400 ≤ 27 distributed load test facility

With mesh strap

Fiber-reinforced shotcrete South African (S.A.)


+ #0-gauge mesh straps distributed load test facility;
220 ≤ 40
behaviour will depend on
pattern of straps

Fiber-reinforced shotcrete South African (S.A.)


+ yielding straps distributed load test facility;
300 ≤ 50
behaviour will depend on
pattern of straps

References
Kaiser, P.K., Tannant, D.D., and McCreath, D.R. 1996. Canadian Rockburst Support Handbook. Geomechanics
Research Centre, Laurentian University, Sudbury, Ontario. p. 314.
Potvin, Y., Wesseloo, J., and Heal, D. 2010. An interpretation of ground support capacity submitted to dynamic
loading. In 5th International Seminar on Deep and High Stress Mining. Edited by M. Van Sint Jan and Y.
Potvin, Santiago, Chile. pp. 251-272.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 253
Stacey, T.R., and Ortlepp, W.D. 2001. Tunnel surface support capacities of various types of wire mesh and
shotcrete under dynamic loading. J. South Afr. Inst. Min. Metall.: 337-342.

Ó 2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Draft manuscript – Copyright protected – Cai and Kaiser 2018


254 Appendix E

Appendix E: Information sheets for rock


support components
2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

E.  Rock support element technical information.................................................................................... 255 


E.1  Cable lacing .............................................................................................................................. 256 
E.2  Cablebolt ................................................................................................................................... 257 
E.3  Chain-link mesh ........................................................................................................................ 258 
E.4  Conebolt .................................................................................................................................... 259 
E.5  D-bolt ........................................................................................................................................ 260 
E.6  Durabar ..................................................................................................................................... 261 
E.7  Duracable .................................................................................................................................. 262 
E.8  Dynamic cablebolt .................................................................................................................... 263 
E.9  Dynamic screen ......................................................................................................................... 264 
E.10  Dynatork - bolt .......................................................................................................................... 265 
E.11  Fiber-reinforced shotcrete ......................................................................................................... 266 
E.12  Friction bolt (Split-set) .............................................................................................................. 267 
E.13  Garford solid bar ....................................................................................................................... 268 
E.14  HEA mesh ................................................................................................................................. 269 
E.15  HE-bolt ...................................................................................................................................... 270 
E.16  MCB conebolt ........................................................................................................................... 271 
E.17  Mechanical bolt ......................................................................................................................... 272 
E.18  Mesh-reinforced shotcrete......................................................................................................... 273 
E.19  Rebar ......................................................................................................................................... 274 
E.20  Roofex ....................................................................................................................................... 275 
E.21  Smooth bar ................................................................................................................................ 276 
E.22  Strap/Mesh strap ....................................................................................................................... 277 
E.23  Swellex bolt .............................................................................................................................. 278 
E.24  TECCO® chain-link mesh ......................................................................................................... 279 
E.25  Threadbar .................................................................................................................................. 280 
E.26  Weld mesh ................................................................................................................................ 281 
E.27  Yield-Loc bolt ........................................................................................................................... 282 
References ............................................................................................................................................ 283 

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 255

E. Rock support element technical information


Many types of rockbolts and surface support elements have been developed and used worldwide. Some of
the rock support elements are reviewed in this appendix. Their technical information is presented in the
form of data sheet, with a brief description of the product followed by discussing the advantages and
disadvantages, major support functions, and how to use it successfully in practice. Typical technical data,
where available, are also presented. The information sheets are presented in alphabetical order.

2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Draft manuscript – Copyright protected – Cai and Kaiser 2018


256 Appendix E

E.1 Cable lacing

Background: Cable lacing was first used in South


2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Africa for rockburst support. This system consists of


grouted rebar and wire mesh, and tensioned steel
cables or laces between rebars in a diamond
configuration. It is used mainly to stabilize broken
rock retained by screen and to dissipate energy. The
Cable
cables act like car seat-belt and can absorb a
significant amount of dynamic energy.
Advantages: Both laboratory tests and field
observations have confirmed that with the addition of
wire rope lacing to the mesh, the capacity of the
support system to absorb energy increases
considerably.
Disadvantages: Labor intensive to install cable
lacing; high cost. These are the reasons that cable
lacing is not widely used in burst-prone mines around
the world.
Major support function: Connect to enhance the
retaining function and energy absorption capacity of
Technical data:
the support system.
Steel designation: High strength steel wire strand
Key to success: Proper installation of cable lacing is
important. Cable lacing, which is very strong, should Cable tensile strength: 1870 MPa
be clamped at every bolt. Otherwise, the lacing may Cable diameter: 12.7 mm
be stretchable and therefore rather “soft,” not stiff Cable length: variable
enough to transfer the baggage load from the
Total displacement capacity: 150 to 400 mm
fractured rocks to the bolts. The combination of weld
Energy absorption capacity: 20 to 40 kJ/m2 (bolts
mesh with lacing performs better than the
and mesh with cable lacing, based on South
combination of chain-link mesh with lacing. Yielding
African drop test result).
bolts is recommended to be used in combination with
cable lacing.
Source: Ortlepp (1997), Stacey and Ortlepp
(2001), Ortlepp and Swart (2002).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 257

E.2 Cablebolt

Background: Cablebolts are used to support stope


hanging walls and backs to enhance stability and
hence reduce dilution. They are also widely used for
support at intersections where rockbolts are deemed
inadequate in length. Most cablebolts used in mining
are plain seven wire strand cablebolts or modified
cablebolts (birdcage, ferruled, nutcase or bulbed
strand) and are often cement grouted. Cablebolts are
usually installed in the second-pass. Spin cablebolts
which use resin grout have been successfully used in
many Canadian mines. The length of spin cablebolts
is limited (< 6 m).
Advantages: It can be installed effectively in very
narrow tunnels; inexpensive; very high load bearing
capacity; high corrosion resistance.
Disadvantages: Tensioning of the cablebolt requires

2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
a special installation procedure; need time for cement
grout to develop strength. Fully grouted cablebolts
have limited deformation and hence energy
capacities. For example, test conducted using the Technical data:
double-embedment dynamic test device on 15.2 mm Steel designation: High strength steel strand
diameter single strand cablebolt, which was cement- Ultimate strength: 1860 MPa
grouted without free length, showed that the cablebolt
Steel diameter: 12.7 mm (0.5”), 15.2 mm (0.6”)
broke at 300 kN after 26 mm displacement (Ortlepp
Yield load: 183 kN (12.7 mm), 260 kN (15.2 mm)
and Erasmus 2005). The energy absorbed was only
3.1 kJ. Length: variable
Major support function: Reinforcement and hold. Steel elongation (min): 3.5%
Key to success: Good grout quality is critical for the Displacement at peak load: 5–10 mm (un-
performance of cablebolt. Cablebolts have a high debonded)
load capacity and de-bonding is required to achieve a Total displacement capacity: 20–40 mm (un-
high deformation capacity. De-bonded cablebolts debonded)
have been used as yielding support elements. For Energy absorption capacity: 3.1 kJ (un-debonded).
example, at New Brunswick Mine, 7 m long The energy absorption capacity of a deboned
cablebolts, de-bonded over a length of 3.7 m, had cablebolt is high, depending on the length of the
been used at intersections along with MCB38 deboned section.
conebolt-based rockburst support system (Falmagne
and Simser 2004). To enhance the holding capacity of
Source: Hoek and Wood (1987), Hutchinson and
the cablebolt, one should consider using cable grip to
Diederichs (1996), Ortlepp and Erasmus (2005),
fasten plates. A large and strong plate will always be
DSI, Mansour Mining Technologies Inc.
beneficial. In fact, it is essential when the holding
function is required.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


258 Appendix E

E.3 Chain‐link mesh

Background: The chain-link mesh is consisted of


individual wire pickets helically wound and
2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

interwoven to form a continuous chain-link with a 50


mm (or 75 mm) nominal diamond mesh. The mesh is
supplied in rolls of varying lengths and widths and
can be galvanized for corrosion protection.
Advantages: Chain-link mesh can accommodate
larger deformation than weld mesh. It is inexpensive,
quick to install, and easy to be repaired; it provides
immediate support after installation. The roll mesh is
suitable for mechanized rock support installation and
results in better surface coverage because there are
less mesh overlaps.
Disadvantages: If a single wire strand fails, the rock
can unravel easily; difficulty to apply shotcrete
through the smaller openings available. Limited load
capacity because small wires (#9 gauge typical) have
to be used to allow roll handling. Lack of sufficient
stiffness to transfer dynamic loads to the bolts
effectively. Technical data:
Major support function: Retain. In addition to the Steel designation: Carbon steel
retaining function, meshes transfer loads to the bolts. Yield strength: 450 MPa
Key to success: In order to fully mobilize the Steel wire diameter (T): variable (#6 and smaller
capacity of yielding bolts, meshes with certain gauge)
stiffness and strength are desired. Previous use of Width: variable
chain-link mesh in burst-prone grounds was based on
Roll length: variable
the consideration of having to ensure sufficient
deformation capacity in the retaining component. Load capacity: 40 kN (#9 gauge, 1.5 m  1.5 m
However, large bulking in the chain-link mesh was sheet)
observed between the MCB conebolts and straps after Displacement at peak load: 420 mm
some rockburst events at New Brunswick Mine in Total displacement capacity: 500 mm
Canada. Hence, the effectiveness of chain-link mesh Energy absorption capacity: < 12 kJ/m2 (based on
to transfer dynamic load is hindered by its softness. static pull test result); < 15 kJ/m2 (based on S.A.
Adding shotcrete, using mesh straps, and reducing drop test result).
bolt spacing are effective methods to increase the
mesh’s stiffness. Use of chain-link mesh in burst-
prone grounds as the main retaining component is not Source: Kaiser et al. (1996), Ortlepp and Stacey
recommended. (1998), DSI.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 259

E.4 Conebolt

Background: The concept and idea of conebolt was


firstly developed in South Africa. The conebolts were
groutable yielding tendons developed by the
Chamber of Mines Research Organization (COMRO)
in 1987 (Jager 1992) for use in cement grouted holes.
When subjected to static loading, the cone (forged
conical head) functions as a wedge-style mechanical
anchor similar to standard mechanical rockbolts.
However, when subjected to dynamic loading, the
conebolts yielded or plough through the grout, thus
absorbing the dynamic energy through a controlled
deformation. To ensure that the cone penetrates
deeply enough into the grout, the cone is flattened on
two parallel sides, allowing the maximum cone
diameter to be increased while keeping the working
area constant. Domed plate and spherical seat are
used to allow better tensile load transfer from the

2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
plate to the bolt. The opposite end to the cone may
have a 'Shepherd's crook' eye to facilitate cable
lacing. Technical data:
Advantages: High energy absorption capacity for Steel designation: NA (not available)
direct impact at the plate, relying on cone sliding for
Yield stress: NA
energy absorption.
Steel diameter: 16 mm, 22 mm
Disadvantages: Need time for the cement grout to
Yield load (average): 100 kN, 200 kN
develop strength; requires a third pass to torque the
bolts once the grout is cured; saponified wax Length: variable
debonding coating could potentially lose its Steel elongation: NA
effectiveness; potential locking-in due to ground Displacement at yield load: 10–20 mm
shear movement; high cost.
Total displacement capacity: > 200 mm
Major support function: Dynamic energy absorption
Energy absorption capacity: >16 kJ (16 mm), 39
and hold. Conebolts are used primarily as the holding
kJ (22 mm).
element in a support system.
Key to success: It is important to use the cement
grout with a proper strength. Good connection to a Source: Ortlepp and Stacey (1994), Kaiser et al.
strong and stiff surface retaining system ensures that (1996), Duraset, CANMET,
the energy can be transferred to the bolt effectively. www.steeledalescs.co.za, Cai et al. (2010).
Use cable lacing or mesh strap to maximize its energy
absorption capacity.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


260 Appendix E

E.5 D‐bolt

Background: D-bolt is a multi-anchored rockbolt


developed in Norway; it is made of a smooth steel bar
2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

with three (or more) anchors along its length, evenly


or unevenly spaced. The bolt is installed using
cement or resin grout. For cement grout, bolts with
either W-anchors or paddle-anchors can be used; for
resin grout, bolts with paddle-anchors are used. Every Paddle-anchor
paddle-anchor consists of two paddles that are formed
by deforming the bolt shank in two orthogonal
directions. The anchorage strength of a paddle-anchor
is greater than the ultimate tensile load capacity of the
bolt; thus, D-bolts mainly absorb energy through
plastic deformation of the steel between anchors.
Advantages: Multiple anchor points ensure reliable
anchorage of the bolt; softer than a fully encapsulated
bolt (e.g. rebar); easy to intall and immediate support Plate & nut
after installation (when resin is used).
Disadvantages: As with other bolts, the thread
section can be the weak link. Recent improvement of
D-bolt made the thread in a section slightly larger Technical data:
than the shaft section, but it requires non-standard Steel designation: NA
nuts and a special dolly for installation. Yield stress (min): 450 MPa
Major support function: Reinforce and hold. The Steel diameter: 20 mm, 22 mm
bolt reinforces the rock by constraining rock dilation
Yield load (min): 190 kN, 230 kN
between anchors, and allows further rock deformation
Length: variable
through the yield of the smooth section between
anchors. Bolt does permit localized opening of Steel elongation: 15%
fracture. Poor bulking control and potential shear Displacement at peak load: depends on the first
failure of the bolt. anchor position.
Key to success: Proper installation is important to Total displacement capacity: depends on stretch
secure the anchorage. Use the right borehole size for length. For 1 m paddel spacing, the displacement
the bolt. It is essential to realize that the energy amount is about 150 mm.
absorption capacity of the bolt relies on steel Energy absorption capacity: 13.1 kJ (20 mm, 0.8
stretching between two anchor points. Hence, when a m stretch length by double-embedment split tube
design requires a dominant dynamic load transfer test), 39 kJ (22 mm, 0.9 m stretch length by
from plate to bolt, energy capacities obtained from double-embedment split tube test).
drop tests with direct impact to the plate are desired;
need case histories to demonstrate the bolt’s
performance in large rockbursts. Source: Dynamic Rock Support, (Li 2010), Li and
Charette (2010), CANMET.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 261

E.6 Durabar

Background: A Durabar is made by forming (cold-


pressed) a sinusoidal wave into a smooth round bar of
ordinary low-carbon steel, acting as a ductile anchor.
The wave shape generates a large but controllable
resistance as the bolt is being pulled out of the hole. Ductile anchor
Adhesion to the cement grout is prevented by a layer
of suitable de-bonding substance that coats the entire
bolt. The “pull-out resistance” depends on the
amplitude and wavelength (pitch) and number of
undulations in the waved portion of the bar. This bolt
was developed in South Africa and used in burst-
prone grounds. De-bonding wax
Advantages: Stiff in static conditions and yield in
dynamic loading conditions; can accommodate large
deformation; easy to install.
Disadvantages: Test data showed that the load

2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
dropped significantly when displacement was greater
than 600 mm (Ortlepp et al., 2001). This could be
attributed to the reduction of sinusoidal wave
amplitudes after sliding over a large distance. Need Technical data:
time for cement grout to develop strength; cannot Steel designation: NA
tension the bolt. Yield stress: 450 MPa
Major support function: Dynamic energy absorption Steel diameter: 16 mm
and hold.
Yield load: 120 kN
Key to success: Use a cement grout with the right
Length: variable
strength because the bolt performance is sensitivity to
Steel elongation: NA
grout strength.
Displacement at peak load: 30 mm
Total displacement capacity: > 600 mm
Energy absorption capacity: 8 kJ / 100 mm
sliding.

Source: Duraset, Ortlepp et al. (2001), Ortlepp


and Swart (2002).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


262 Appendix E

E.7 Duracable

Background: A Duracable is made of a cable that is


sheathed with a length of close-fitting steel tubing of
2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

2.5 mm wall thickness to form a sleeve over the Cablebolt


anchoring length of the unit. The sinusoidal waves
are cold-pressed into the portion of the cable that is
surrounded by the steel sleeve, and the deformed
sleeve is stiff enough to retain the crinkled shape in Steel tube
the cable. When installed, the deformed sleeved end
of the Duracable is anchored at the remote end of the
hole with the entire remaining “free” length left un-
grouted or de-bonded by PVC sleeve in fully-grouted
holes. When the cablebolt is loaded dynamically,
controlled sliding will occur in the crinkled portion Debonding agent
before the transient load can reach values close to the
breaking load of the cable. Barrel and wedge
Advantages: Yielding device with large displacement
capacity.
Disadvantages: Load could drop after the device
starts to slide.
Technical data:
Major support function: Dynamic energy absorption
and hold. Steel designation: High strength steel strand

Key to success: A minimum of grout strength of 25 Ultimate strength: 1860 MPa


MPa is required. Plating is important to utilize the Steel diameter: NA
holding capacity of the cablebolt. Further tests are Yield load (min): 100 kN (static sliding)
needed to characterize the mechanical properties of
Length: variable
the cablebolt. Need case histories to demonstrate the
Steel elongation: %
cablebolt’s performance in real rockburst conditions.
Displacement at peak load: 25–30 mm
Total displacement capacity: > 220 mm
Energy absorption capacity: > 19.3 kJ (the
Duracable was intact after 220 mm displacement).

Source: Duraset, Ortlepp and Erasmus (2005).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 263

E.8 Dynamic cablebolt

Background: A dynamic cablebolt, which uses seven


wire 15.2 mm compact strand, has a yielding
mechanism attached at one end of the debonded
compact strand and a barrel, plate and wedge at the
end to provide a point loaded yielding anchor. The
king wire is removed from the cable and replaced
with a larger 8 mm diameter spring wire. It works in
principle the same as the Garford solid bar, wherein Debonding sleeve
the cable pulls through the yielding device enabling
the cable to absorb energy and remain intact. It is
used in areas where seismic activity is expected.
Advantages: Consistent yielding load due to the Yielding device
yielding device used; large energy absorption
capacity.
Disadvantages: The cablebolt is installed using
cement grout. Hence, time is needed for the grout to

2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
develop strength.
Major support function: Dynamic energy absorption
and hold. For example, 6 m long dynamic cablebolts
Technical data:
(300 mm travel and 14 ton slip rating) on 1.5 m
spacings and 1.2 to 2 m ring spacings were used Steel designation: Seven Wire Compact Strand
along with mesh straps (5.6 mm wire) and DE plates Ultimate strength: 1860 MPa
(300 mm circular rubber backing plate) as the Steel diameter: 15.2 mm
secondary support at Junction Mine in Australia (Li
Yield load (min): 212 kN (steel), 150 kN (static,
et al. 2003; Li et al. 2004) . In one case, in which the
device yielding), 80–120 kN (dynamic, device
dynamic support system survived a large rockburst,
yielding)
the drive convergence was up to 0.7 m over a 20 m
Length: variable
length of the drift from the stope brow. The
fragmented rock around the excavation was fully Steel elongation: 3.5%
contained. The dynamic support system has Total displacement capacity: up to 300 mm
demonstrated its capability of surviving moderate Energy absorption capacity: 30 kJ (at 300 mm
rockbursts. displacement).
Key to success: Good grout quality is critical for the
performance of cablebolts; need case histories to
Source: Garford Pty Ltd.
demonstrate the cablebolt’s performance in large
rockbursts. Plating is important to utilize the holding
capacity of the cablebolt.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


264 Appendix E

E.9 Dynamic screen

Background: The dynamic screen was developed


based on the consideration of the benefit of using 0
2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

gauge (0.305”) mesh straps to enhance the capacity of


the surface support system while ensuring easy
installation. The 0 gauge wires are integrated into the
conventional wire mesh for one-pass rock support
system installation. The top screen shown on the left
is a combination of standard welded wire mesh with 4
leading strands made of 0 gauge wire. The bottom
screen is a combination of standard welded wire
mesh with 4 leading edge strands and 4 central
strands made of 0 gauge wire for additional strength
enhancement. The dynamic screen can be used with
dynamic support or in combination with any other
type of ground support.
Advantages: Less parts to handle; improved
productivity; improved surface support capacity.
Disadvantages: High cost.
Major support function: Retain and connect. The
Technical data:
integrated 0-gauge wires fullfil the connection
function of traditional mesh straps. Steel designation: Carbon mild steel

Key to success: Proper overlap between screens; need Yield strength: NA


test data to show the capacities of the mesh; need Steel wire diameter: variable (#0 strands and the
case histories to demonstrate the mesh’s performance base wires (#4 and #6 are most commonly used))
in large rockbursts. Sheet size: variable (available in all widths and
lengths)
Load capacity: NA
Displacement at peak load: NA
Total displacement capacity: NA
Energy absorption capacity: NA

Source: Mansour Mining Technologies Inc. (2014).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 265

E.10 Dynatork ‐ bolt

Background: Dynatork - bolt is a new dynamic bolt


designed with a spiral mixing blade to ensure
ultimate mixing of resin in the borehole. The bolt
diameter is 19.05 mm and the mixing blade is
attached to the smooth bar. The blade is used to mix
the resin properly and to provide anchorage for static
ground support. There is an offset between the blade
and the smooth bar, which allows the bolt to yield
and transfer dynamic loads into the resin in rockburst
conditions, absorbing the energy through controlled
deformation. This ploughing effect through the resin
can ensure that most of the kinetic energy in a
rockburst is contained within the rockmass.
Advantages: Good resin mixing; tested by direct
impact at the plate; immediate support after
installation.

2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Disadvantages: Potential locking-in due to ground
shear movement.
Major support function: Dynamic energy absorption
Technical data:
and hold.
Steel designation: 1055M
Key to success: Need case histories to demonstrate
the bolt’s performance in large rockbursts. Yield stress (typical): 560 MPa
Steel diameter: 19.05 mm
Yield load (min): 93 kN
Ultimate tensile load: 143 kN
Length: variable
Steel elongation: 10%
Displacement at peak load: 138–380 mm
Total displacement capacity: 180–300 mm
Energy absorption capacity: 16–30 kJ. 16 kJ at
180 mm displacement; 23 kJ at 270 mm
displacement; 30 kJ at 294 mm displacement.

Source: Oler (2012).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


266 Appendix E

E.11 Fiber‐reinforced shotcrete

Background: Shotcrete, as an effective surface


support element, has been used in rock support for
2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

many years. Plain shotcrete is too brittle and fiber or


mesh reinforcements are needed to enhance its load
and deformation capability. Reinforcement fibers can
be in the form of steel or synthetic fiber. Shotcreting
the walls right after excavation can maintain the
integrity of rock masses and hence enhance the
stability of the opening. In addition, shotcrete is
frequently used to rehab damaged drifts.
Advantages: The use of fiber-reinforced shotcrete
brings the benefit of minimizing the labor-intensive
process of mesh installation. It allows the exact
contours of the rock with the surface support. Fiber-
reinforced shotcrete is stronger and has better crack
resistance than plain shotcrete.
Disadvantages: Limited deformation capacity. Static
tests performed by Kirsten and Labrum (1990) shows
that fiber-reinforced shotcrete performs better than
mesh-reinforced shotcrete when deformations are Technical data:
relatively small and the shotcrete is uniformly Shotcrete strength: 25–35 MPa (28 day strength)
strained. Fiber-reinforced shotcrete panel on its own Fiber: steel, synthetic
is not suitable for concentrated or repeated dynamic
Thickness: 50–75 mm (most common)
loading.
Peak load: 20–30 kN
Major support function: Retain. Shotcrete also
Displacement at peak load: 2–20 mm
plays a role to keep the integrity of a rock mass and
hence enhance its strength. Energy absorption capacity: 3–6kJ/m2 (direct
Key to success: Fiber-reinforced shotcrete alone impact to panel at a 1.2 m  1.2 m diamond
cannot provide effective surface support under bolting pattern).
rockburst conditions. A good practice is to install
mesh with rockbolts over fiber-reinforced shotcrete to Source: Kirsten and Labrum (1990), Kaiser et al.
form an integrated system. This allows distributed (1996), Tannant (1997), Morton et al. (2009).
load to be transferred to the holding elements for
effective energy absorption and deformation control.
In drifts that may experience large ground
movements, whether it is due to rock bulking or
squeezing, shotcrete panels instead of closed
shotcrete rings should be used.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 267

E.12 Friction bolt (Split‐set)

Background: Friction bolts, also called Split-set


stabilizers, friction sets, friction stabilizers, consist of
a slotted high strength steel tube and a faceplate
(Scott 1977). It is installed by pushing it into a
slightly undersized-hole and the radial spring force
generated, by the compression of the C shaped tube,
provides the frictional anchorage along the entire
length of the hole. Grouting the pipe using cement
can be conducted to increase its load capacity. A
grouted FS46 bolt with 1 m of embedment can hold a
rock weighing 10 to 15 tonnes (Villaescusa and
Wright 1997). Grouted friction bolts also have a
higher shear load capacity. It should be noted that
grouting of friction bolts leads to stiff bolt response.
Advantages: Easy installation (suitable for
mechanization); inexpensive and provide immediate

2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
support when installed. They are particularly useful
in fractured rocks where securing a point anchor or
inserting resin into boreholes is difficult. It is also
useful in mild bursting ground because it can Technical data:
accommodate large deformation. For example, FS46 Steel designation: High strength steel
friction bolts with #4 weld mesh were successfully
Tube diameter: 33 mm (FS33), 39 mm (FS39),
used at Vale’s Creighton Mine and Copper Cliff
and 46 mm (FS46)
Mine in Canada as a dynamic support system
Sliding load: about one ton (= 0.907 tonnes) per
(Punkkinen and Mamidi 2010).
foot length (FS46), or 3 ton/m.
Disadvantages: The bolts are susceptible to
Length: variable
corrosion, and hence its long-term performance is
poor. However, the bolt can be galvanized to provide Displacement at peak load: 10–30 mm
some protection against corrosion. Anchorage Total displacement capacity: > 100 mm
capacity is sensitive to borehole size and rock Energy absorption capacity: 2.71.0 kJ per 100
condition. Because of their unreliable anchorage mm of bolt sliding (FS46), which is generally lower
capacity, friction bolts are not used for roof support than its static energy absorption capacity of 3 to 6
in most deep mines in Canada. kJ per 100 mm of sliding. For a 200 mm sliding
Major support function: Hold. The holding capacity distance (wall deformation) the dynamic energy
may be reduced if the shallow skin of the rocks is capacity is about 4 to 7 kJ.
highly fractured. The capacity is proportional to the
anchor length.
Source: Mansour Mining Technologies Inc., DSI,
Key to success: A borehole needs to be the right size. Hoek and Wood (1987), Ortlepp and Stacey
A large hole reduces the holding capacity while a (1999), Player et al. (2009b).
smaller hole can damage the bolt as it is pushed into
the hole. Hence, it is important to use the right bit to
drill the borehole.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


268 Appendix E

E.13 Garford solid bar

Background: A Garford solid bar (or Garford


dynamic bolt) is made of an anchor and a coarse
2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

threaded steel sleeve crimped on to the end of the bolt


as the resin mixing device. The remainder of the bolt
Mixing device
is covered in a polyethylene sleeve. The engineered
anchor (or the yielding device) allows the bolt to
stretch by an extruding process, and the diameter of Debonding sleeve
the solid bar is reduced from its original size to a
smaller one at the position of the anchor. The
extruding force remains approximately constant. An
end stop mechanism allows the bolt to lock when the
pre-determined displacement capacity is exhausted. Yielding device
The bolt can be anchored in the borehole using
cement as well.
Advantages: Consistent load due to the extruding Plate & nut
mechanism; immediate support after installation
when resin is used; superb energy absorption capacity
for direct impact at the plate.
Disadvantages: The extruding length is pre-
determined (up to 500 mm); longer boreholes are Technical data:
required to accommodate the extruding length; need Steel designation: 5152AVH grade steel
large diameter boreholes; high cost. Yield stress (min): 550 MPa
Major support function: Dynamic energy absorption Steel diameter: 21.7 mm
and hold.
Yield load (min): 199 kN
Key to success: Proper bolt installation and
Length: variable
particularly effective mixing of the resin capsule
Steel elongation: 12%
components in large diameter boreholes (typically 45
mm) are critical; need case histories to demonstrate Displacement at peak load: 3–4 mm
the bolt’s performance in large rockbursts. Total displacement capacity: up to 500 mm
Dynamic load capacity: 100–125 kN
Energy absorption capacity: 27–33 kJ (impact at
the plate).

Source: Garford Pty Ltd, Player et al. (2008),


Varden et al. (2008).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 269

E.14 HEA mesh

Background: HEA (High Energy Absorption) mesh


was developed at the University of Western Rockbolt & plate
Australia, out of the consideration to retain the
benefit of adding cable lacing to the surface support
system while ensuring easy installation. Cable lacing
is pre-fabricated over a commonly used 2.4  3 m2
sheet of mesh (5.6 mm wire, 100  100 mm2
aperture), which can be either standard or crinkled.
This product fails to achieve the goals of
conventional cable lasing which are: use of long
strechable cable runs that add ductility at unknown
locations between cable anchor points. Weld mesh
Advantages: Provide immediate support after
installation. Quick installation compared with
traditional cable lacing procedures. Cable lacing
Disadvantages: In actual cable lacing installation, all

2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
lacing ropes are connected to the holding bolts in a
secured fashion. The cable lacing in the HEA mesh
however is connected to the bolts through plates, and Technical data:
this connection is weaker than direct connection of
Mesh wire diameter: 5.6 mm
cables to the bolts. If one bolt fails, the lacing system
will be affected. Potential bend loading to bolts if the Wire yield strength: 450 MPa
cables take load may snap bolt heads. High cost and Cable diameter: 12.7 mm
manual mesh handling is difficult due to heavy Cable tensile strength: 1870 kN
weight. The pre-defined lacing pattern requires
Sheet size: 2.4 m  3 m
borehole drilling yhat will result in little flexibility
for rockbolt installation. Mass per sheet: 45.5 kg
Major support function: Retain and connect; load Peak load: 170 kN if applied at middle of a a 2.4
transfer from surface support to holding support m  3.0 m sheet (providing about 30 kN/m2)
elements. Total displacement capacity: 900 mm for load
Key to success: applied at middle of a 2.4 m  3.0 m sheet
Correct overlay of HEA mesh sheets and bolt Energy absorption capacity: 30 and 40 kJ for
placement is critical to ensure cable lacing performs approximate 5.5 m2 (estimated based on static pull
as a complete system. Having mechanized rock test result)
support equipment is critical for the use of this
product. Source: Potvin (2009), Potvin and Heal (2010).
WARNING: Due to the shear loading of the
boltheads by the lacing, the authors do not
recommend use of this mesh in burst-prone ground
until case histories have been provided to
demonstrate the mesh’s effectiveness in severe
rockburst conditions.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


270 Appendix E

E.15 HE‐bolt

Background: HE-bolt was developed by Prof.


Manchao He (Patent No. ZL 2010 1 0196197.2) at
2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

the State Key Laboratory (SKL) of Geomechanics &


Deep Underground Engineering (GDUE) in China
University of Mining & Technology Beijing Pallet
(CUMTB). It is a compound device consisting of a Cone
Nut
Sleeve
piston-like cone bonded tightly to a bolt shank sliding Bolt shank
in an elastically deformable sleeve pipe. The working
Anchored end
resistance of the bolt is actually the frictional force
generated by the relative sliding between the cone
and the sleeve. Thus, the load resistance is dependent
Elongation length
on the structural parameters of the cone and elasticity
of the metal sleeve. The elongation of the bolt is
realized by the relative displacement of the cone-
shank assembly with respect to the sleeve pipe.
Advantages: Large deformation capacity (elongation
up to 1000 mm) at high constant resistance; high
energy absorption capacity (up to 200 kJ); proven
support performance in weak rock and rockburst
conditions. Technical data:
Disadvantages: Water-bearing rock condition would Steel designation: NA
impair the constant-resistant performance; high cost. Yield stress: NA
Major support function: Containing large Steel diameter: 20 mm, 22 mm
deformation and absorbing energy impact in deep
Sliding load (min.): 130 kN (HE-bolt13), 160 kN
ground conditions.
(HE-bolt20)
Key to success: Diameter of the drilled hole in the
Length: variable
rock mass should be 2 to 3 mm larger than the outer
Steel elongation: NA
diameter of the sleeve pipe; the resin should be
sufficiently stirred; pre-tightening tensile force should Displacement at peak load: NA
be applied using air-powered hexagon spanner with a Total displacement capacity: up to 1000 mm.
hexagon box-ended wrench suited for the hexagon Energy absorption capacity: 130 kJ (HE-bolt13),
nut on the HE-bolt. Dynamic tests conducted so far 160 kJ (HE-bolt20) based on static test results with
were limited to low single energy input (9.8 kJ). a 1000 mm sliding range.
Additional dynamic tests need to be conducted to
reveal the true dynamic capacity of the bolt.
Source: State Key Laboratory (SKL) of for
Geomechanics & Deep Underground Engineering
(GDUE) in China University of Mining &
Technology Beijing (CUMTB); He et al. (2012);
He et al. (2014).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 271

E.16 MCB conebolt

Background: The South African conebolts were


modified to accommodate resin-grouting applications
Resin mixing blade Dome plate
in Canada. Norand Inc. modified the conebolt head Debonding agent
and added a blade for the purpose of resin mixing
and the new bolt was called Modified Cone Bolt
(MCB). The blade length was extended by Mansour
Mining Technologies Inc. to facilitate better resin Cone Spherical seat & dome nut
mixing and torque resistance during bolt tightening.
Grease is used as the debonding agent for MCB38, MCB38
while PVC sleeve is used as the debonding agent for
MCB33. MCB33 and MCB38 conebolts are installed
in 33 and 38 mm diameter boreholes, respectively.
Advantages: High and consistent energy absorption
capacity (MCB33), tested by direct impact at the
plate; immediate support after installation; no grease
in MCB33 design. MCB33

2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Disadvantages: Grease debonding coating (MCB38)
could potentially lose its effectiveness; potential
locking-in due to ground shear movement; high cost.
Technical data:
Major support function: Dynamic energy absorption
and hold. Steel designation: C1055M

Key to success: Need to use the right resin size for Yield stress: 448 MPa
MCB38 and MCB33 respectively; use resin with a Steel diameter: 17.2 mm
proper strength. Proper resin mixing is important to Yield load (min): 98.5 kN (at thread)
ensure that the dynamic response of the MCB
Length: variable
conebolts is consistent on a repetitive basis as
Borehole diameter: 38 mm (MCB38), 33 mm
designed. For proper bolt installation, spin the bolt at
(MCB33)
full rotational speed while at the same time
advancing the bolt through the resin, and then Steel elongation: 10%
continue spinning for additional 30-40 revolutions to Displacement at yield load: 20–40 mm (MCB33)
complete the mixing process. Total displacement capacity: 300–900 mm
MCB conebolts are recommended for use with #0 (MCB33)
gauge mesh straps. MCB33 uses standard 33 mm Energy absorption capacity: 30–35 kJ (MCB33).
drill bit; hence, it is possible to install rockburst Impact at the plate (direct loading).
support in a one-pass system. Field experience shows
that it is difficult to install second-pass bolts in
ground that was damaged by previous seismic Source: Mansour Mining Technologies Inc., Cai
activity or ground movement. Therefore, a one-pass and Champaigne (2009, 2012), Cai et al. (2010),
system provides an option to enhance productivity CANMET.
and safety.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


272 Appendix E

E.17 Mechanical bolt

Background: A mechanical rockbolts generally


consists of a plain steel rod with a mechanical anchor
2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

at one end and a faceplate and nut at the other. The


expansion shell mechanical bolt is the most
commonly used type. Mechanical bolts are usually
tensioned to about 70% of the ultimate breaking load
or 50% of the yield load after installation. An
approximately linear relationship exists between the
applied torque and the tension in the bolt.
Advantages: Quick to install; inexpensive and can
provide immediate support after installation.
Disadvantages: Difficult to install reliably (need the
correct amount of torque); subject to corrosion if not
post-grouted; rockbolts could lose their tension due to
anchor slip over time, vibration caused by blasting or
seismic waves, or rust caused by aggressive
groundwater. It had been observed that some
mechanical bolts were projected along with the
fractured rocks in some rockbursts. Field tests at
Compbell Mine in Canada showed that rock support Technical data:
systems using mechanical bolts performed the worst Steel designation: C1055, C1070
compared with other bolts. Hence, many hard rock Yield stress (min): 348 MPa (C1055), 383 MPa
mines in Canada had stopped using mechanical bolts (C1070)
in burst-prone grounds.
Steel diameter: 16 mm (5/8”), 19 mm (3/4”), 22
Major support function: Reinforce and hold. mm (7/8”),
Mechanical rockbolts are mainly used as temporary
Yield load (min): 69 kN (5/8”, C1055)
support, but can be post-grouted to provide
Length: variable
“permanent” support. A modified mechanical bolt is
CT-Bolt (www.ctbolt.com), which has a special Steel elongation (min): 10%
hemispherical dome at the end which serves both as a Displacement at peak load: 10–40 mm
holder of the rock loading on the plate and as a Total displacement capacity: 20–50 mm
grouting chamber.
Energy absorption capacity: between 0.9 and 5 kJ
Key to success: The rock must be hard enough to (F 1 ¼ shell).
provide a good grip for the mechanical anchor. The
diameter of the hole must be carefully controlled,
because an oversized hole can result in poor Source: Hoek and Wood (1987), Mansour Mining
anchorage, and an undersized hole will not permit Technologies Inc., DSI, CANMET.
installing the expansion shell properly. The hole
should be free of small rock pieces to ensure the bolt
can be inserted smoothly. Grouting may prevent
potential projection of the bolt in a rockburst event.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 273

E.18 Mesh‐reinforced shotcrete

Background: Mesh is relatively soft and provides


very little support at small wall deformations. Adding
shotcrete to the mesh, i.e., forming mesh-reinforced
shotcrete solves the problem of initial low stiffness; it
also increases the peak load capacity of the support
element. In addition, the mesh retains its ductile
capacity and mesh-reinforced shotcrete can have
almost the same deformation capacity as the mesh
acting alone. Combined with yielding bolts, mesh-
reinforced shotcrete can dissipate a large amount of
dynamic energy during rock fracturing. Mesh-
reinforced shotcrete is nicknamed “supermesh,” and
it is a good example of combining stiffness and
softness in a rock support system.
Advantages: Shotcrete protects the mesh from
corrosion and prevents "unzipping" arising from

2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
individual broken wires or from inadequate mesh
overlaps. Shotcrete also spreads the load from
individual rock blocks evenly and improves the
connection to holding elements. Mesh-reinforced Technical data:
shotcrete has favorable stiffness to transfer load to Shotcrete strength: 25–35 MPa
yielding bolts. Mesh-reinforced shotcrete performs
Mesh: #4, #6
better when large deformations causing extension
Thickness: 50–75 mm (most common)
cracking are encountered and localized cracking due
to strain localization occurs. Peak load: 45–55 kN (#6 weld mesh)
Disadvantages: Need a two-pass installation Displacement at peak load: 30–50 mm
procedure. It requires filling in all the voids behind Total displacement capacity: 100–250 mm
the mesh with shotcrete, more than what is needed by Energy absorption capacity: 15–20 kJ/m2 (direct
the minimum specified thickness of shotcrete; impact to panel at a 1.2 m  1.2 m diamond
sometimes, mesh cannot be fully covered; higher bolting pattern).
rebound for shotcrete with larger aggregates;
shotcrete pieces may be ejected in a rockburst if a
large amount of shotcrete is spayed over mesh Source: Kaiser et al. (1996).
strands; high cost.
Major support function: Retain and dynamic load
transfer.
Key to success: In drifts that may experience large
ground movements, whether it is due to rock bulking
or squeezing, shotcrete panels instead of closed
shotcrete rings should be used. Yielding support
system (e.g., yielding bolts with straps) installed on
top of the mesh-reinforced shotcrete is preferred.

 
Draft manuscript – Copyright protected – Cai and Kaiser 2018
274 Appendix E

E.19 Rebar

Background: Dowels or anchor bars or rebars,


generally consist of deformed steel bars, are grouted
2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

into the rock. The deformed (or ribbed) bar serves the
purposes of better mixing the catalyst and resin, and
increasing frictional grip between the bar and the
grout. Grouting material can be cement or resin. A
faceplate and nut are often used along with the rebar.
Rebar is one of the most widely used rockbolts in
rock support. Thread
Advantages: Simple and inexpensive; high load
capacity and immediate support when resin is used;
fully grouted rebars are highly resistant to corrosion.
Disadvantages: Stiff support; limited deformation
capacity. Faceplate
Major support function: Reinforcement. Rebars
reinforce the rock mass and prevent the rocks near the Pin nut
excavation boundary from fracturing into smaller
pieces. It reduces the bulking factor. This
reinforcement action is very critical for the rock –
Technical data:
rock support interaction because it allows the
dynamic load to be transferred effectively to the Steel designation: Grade 60
yielding bolts. Rebars can be used for both temporary Yield stress: 414 MPa
and permanent support. Steel diameter: 19 mm (3/4”), 22 mm (7/8”), 25
Key to success: Proper resin mixing is critical for the mm (1”)
bolt to achieve satisfactory performance; using the Yield load (min): 120 kN (3/4”), 155 kN (7/8”), 200
right hole size and resin cartridge as well as drilling kN (1”, forged head); 89 kN, 123 kN, 162 kN
the right hole length (50 mm longer than the bolt (threaded head)
length) is important to ensure full encapsulation of
Length: variable
the bolt; install the bolt before significant rock mass
Steel elongation: 9–13%
deformation has occurred.
Displacement at peak load: 5–10 mm (threaded
Rebars are generally not tensioned right after
head)
installation. However, when tensioning is required, a
technique employing quick-set and slow-set resins Total displacement capacity: 5–30 mm (threaded
can be used. Quick-set resin is placed to the far end of head). Can be larger if grout quality is
the hole. As the quick-set resin hardens, the rebar can compromized by stress fracturing near collar.
be tensioned. Another method for tensioning rebars is Energy absorption capacity: < 6 kJ when failed at
to use a rebar with a mechanical anchor attached at thread (impact at the plate); 12–14 kJ (22 mm)
the far end (e.g., InSta’l bolt from Jennmar). when failed across a joint with 50 cm free
Rebars should be installed as soon as possible after stetching length (double-embedment split tube
excavation to fulfill their major role as rock test).
reinforcement units.
Source: Mansour Mining Technologies Inc., DSI,
CANMET, Simser (2007), Labrie et al. (2008).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 275

E.20 Roofex

Background: Roofex is a new yielding support


tendon developed at Atlas Copco Inc. It consists of a
fixed element inside which a smooth bar can slide Mixing / stop element
and generate a constant frictional resistance. Because
the sliding occurs at metal-to-metal contacts, by a
mechanism of cut through energy absorber with pins, gth
g l en
the frictional resistance is fairly consistent. If the Energy absorber i di n
Sl
sliding displacement capacity is consumed, the bolt
will be locked up when the resin mixing device
reaches the energy absorber shell and the load will
increase and may cause the steel to yield. Steel pins
and carbide pins are used for Roofex bolts intended
for static and dynamic supports, respectively.
Advantages: Consistent load due to the designed Face plate and Nut
sliding mechanism; high energy absorption capacity
for direct impact at the plate.

2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Disadvantages: The displacement capacity of the
bolt depends on the sliding length of the bolt, which
has to be pre-defined (varying from anywhere
between 20 to 100 cm, and the practical length is Technical data:
between 20 and 60 cm). Longer boreholes are Steel designation: CK 45
required to accommodate the sliding length; steel Yield stress: 733 MPa
stretching increases with impact energy; very high
Steel diameter: 12.5 mm (R8D), 20 mm (R20D,
cost. Unfortunately, Atlas Copco decided in 2013 to
prototype as of 2010)
stop selling Roofex bolts.
Sliding load: 80 kN (R8D), 200 kN (R20D)
Major support function: Dynamic energy absorption
and hold. Length: variable
Key to success: Proper resin mixing is critical for Steel elongation: 10%
securing the anchor (energy absorber); need case Displacement at peak load: 5–10 mm (R8D)
histories to demonstrate the bolt’s performance in Total displacement capacity: > 200 mm (depends
large rockbursts. on the pre-defined sliding length)
Energy absorption capacity: 6.73 kJ + 3.03 kJ/per
Unfortunately, this product is no longer available in 100 mm plate deformation (R8D); 7.52 kJ + 5.39
the market place. kJ/100 mm plate deformation (R20D). Impact at
the plate.

Source: Atlas Copco, CANMET, Plouffe et al.


(2008), Doucet and Gradnik (2010).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


276 Appendix E

E.21 Smooth bar

Background: A smooth bar, as the name indicates, is


a smooth bar with threads at the collar end only. The
2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

bolt is installed into a borehole using cement grout. It


relies on the friction between the bar and the grout to
provide support. Although it is not widely used for
rock support, its potential as a yielding bolt had been
demonstrated through laboratory and field tests. It is Smooth bar
shown from a blasting test of reinforced concrete
blocks anchored to a rock floor that 16 mm diameter
smooth bars performed almost as well as 16 mm
diameter conebolts, arresting the dynamic energy of
the ejected concrete block without bolt damage
(Ortlepp 1994).
Advantages: Simple to make and inexpensive. Face plate and Nut
Disadvantages: Need time for the grout to develop
strength; cannot tension the bolt. Potentially low
anchorage force when not properly grouted.
Major support function: Hold and reinforce.
Key to success: The grout cannot shrink in volume Technical data:
when hardens. Need case histories to demonstrate the
Steel designation: NA
bolt’s performance in large rockbursts.
Yield stress: NA
Steel diameter: 16 mm
Yield load (min): NA
Length: variable
Steel elongation: NA
Displacement at peak load: 4 mm
Total displacement capacity: > 200 mm
Energy absorption capacity: (3.0L-0.15) kJ/100
mm sliding, where L is the initial embedded length
in meters.

Source: Ortlepp (1994), Ortlepp and Stacey


(1999), Player (2004).

 
Draft manuscript – Copyright protected – Cai and Kaiser 2018
Rockburst Support Reference Book (I) 277

E.22 Strap/Mesh strap

Background: Straps are used extensively in mining


and civil engineering applications to hold slabby
ground between rockbolts or to prevent slabs from
loosening. Field observation and dynamic test results
have shown that adding metal or mesh straps to the
support system is an effective means to increase the
system’s overall load and energy absorption
capacities. Mesh strap
Rockbolt straps, which are made from steel sheets,
are strong and stiff and hence their installation can be
Yielding bolt
difficult. On the other hand, mesh straps are relatively
flexible and easy to install. The dimension of
commonly used mesh straps is 30 cm in width, 10 cm
in wire aperture. The mesh straps are made from 8.23
mm diameter wires. Durastraps are made by four 8
mm diameter wires transversely linked by 6 mm

2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
diameter mild steel wires, ‘pig-tail’ wrapped around
the 8 mm wires at intervals.
Advantages: Increase the surface support load and
energy absorption capacities. It can prevent the bolts Technical data:
from damaging mesh under the mesh straps. Straps Steel designation: mild steel
are easy to install with rockbolts. Compared with Strap type: metal strap, mesh strap
cable lacing, straps are relatively stiffer, cover a large
Strap length: variable
area, and can transfer dynamic load to the bolts more
Total displacement capacity: 120–200 mm
effectively than cable lacing.
Energy absorption capacity: 42–55 kJ/m2 (bolts,
Disadvantages: Additional effort is required for
mesh with mesh strap, based on S.A. drop test
installation. Need a second-pass.
results. Note: this energy capacity is the enhanced
Major support function: Connect to enhance the
system capacity with mesh strap; it is not the
retaining function and energy absorption capacity of
energy capacity of the mesh strap alone).
the support system.
Key to success: Straps should always be installed
Source: Ortlepp (1997), Stacey and Ortlepp
over mesh or shotcrete to maximize the system’s
(2001), Ortlepp and Swart (2002).
capacity. Yielding bolts should be used in
combination with straps in burst-prone grounds.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


278 Appendix E

E.23 Swellex bolt

Background: A Swellex bolt (Atlas Copco, 1982) is a


folded tube that has an undersized profile and when
2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

inserted into a borehole, it is inflated by high pressure


water (up to 30 MPa) to partially contour the bolt to
the profile of the borehole, thus providing frictional
resistance. Expansion of the deformed tube leads to a
reduction in bolt length, which puts the bolt into
tension and tight the plate against the rock surface.
Advantages: The bolt can be installed quickly and
provide immediate support after installation; suitable
for mechanized rock bolting. It is especially useful in
fractured or altered grounds where installation of
resin bolts or friction bolts are difficult. In its original
state, the diameter of a Swellex bolt much smaller
than the hole diameter, thus making it easier to
install. Because the Swellex bolt can conform to the
borehole wall profile, its frictional resistance per unit
length can be higher than that of the friction bolt. The
diameter of the drill hole is not as critical as other
types of rockbolts such as friction bolts and Technical data:
mechanical bolts. Steel designation: S355 JR
Disadvantages: Relatively high cost and Yield stress: 355 MPa
susceptibility to corrosion; weak heads, which means
Steel diameter: Mn12, Mn16, Mn24
that if the impact energy is applied to the face plate,
Break load (min): 110 kN, 160 kN, 200 kN
the energy absorption level will be somewhat smaller
(< 1 kJ for direct impact loading at the face plate). Length: variable
The bolt becomes stiff when the ground moves. Steel elongation: 20%
Major support function: Hold. If the ground is Displacement at peak load: 20–40 mm
deformed, the frictional resistance of the bolt can Total displacement capacity: > 80 mm at plate
increase significantly, rendering it somewhat stiff.
Energy absorption capacity: 6.21.3 kJ per 100
Hence, the energy capacity of the bolt may be
mm of bolt sliding (Mn24). A NTC test on a 2.1 m
reduced.
long Mn12 Swellex showed that on a single impact,
Key to success: Proper installation with the right the bolt can dissipate over 9 kJ of energy (data
inflation pressure. Use plastic or zinc rich resin epoxy based on double-embedment split tube testing
primer coated bolts in corrosive environment. method).

Source: Atlas Copco, NTC (Noranda Technolgy


Center), CANMET, Hoek and Wood (1987),
Ortlepp and Stacey (1999), Player et al. (2009b).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 279

E.24 TECCO® chain‐link mesh

Background: TECCO® chain-link mesh is a product


originally developed for rockfall protection in high
slopes. The wires at the selvedge are bent over and
double twisted in such a way that this connection is
as strong as the mesh itself. The high tensile strength
wire used is about 4 times stronger than the one used
to make conventional meshes. Hence, the load
capacity of a TECCO® chain-link mesh is 4 to 5
times higher than a conventional chain-link mesh.
The mesh is supplied in rolls of 30 m × 3.5 m, and it
has an aluminum-zinc coating, providing a higher
corrosion resistance than standard galvanized coating.
Advantages: High load and energy absorption Spike plate
capacities; difficult to break at plate connections (an
interesting observation from a laboratory pull test was
that the high tensile strength mesh wire actually cut a
TECCO mesh

2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
standard grade steel plate). Suitable for mechanized
rock support installation; better surface coverage due
to less mesh overlaps compared with weld mesh.
Disadvantages: High cost. Lack of sufficient Technical data:
stiffness to transfer dynamic loads to the bolts Steel designation: High strength steel wire
effectively. Yield strength: 1770 MPa
Major support function: Retain. Steel diameter: 3 mm, 4 mm
Key to success: It is recommended to use special Wire breaking load: 12 kN (3 mm), 22 kN (4 mm)
spike plates for TECCO® chain-link mesh because
Mesh load capacity: 190 kN (1 m × 1 m bolt
the load can be transferred to the plates from more
pattern, 4 mm)
than one wire. Furthermore, connection clips can be
Total displacement capacity: > 300 mm
used to better link adjacent mesh sheets. Need case
histories to demonstrate the mesh’s performance in Energy absorption capacity: 50 kJ/m2 (estimated,
large rockbursts. 4 mm. Note: this energy capacity is the enhanced
system capacity with the high strength mesh; it is
not the energy capacity of the high strength mesh
alone).

Source: Geobrugg, Roth et al. (2004), Bucher et al.


(2010).

 
Draft manuscript – Copyright protected – Cai and Kaiser 2018
280 References

E.25 Threadbar

Background: Rebars have a cut down section for


manufacturing the thread, and the threaded section
often becomes the weakest link in the bolt.
2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Threadbars have external hot-rolled coarse threading


(10 mm pitch) over the whole length of the bar,
which can avoid the weak section by conventional Threadbar
threading. This will allow the bar to cope with a
relatively higher level of elongation than rebars.
Another feature is that couplers can be used to extend
the bolt length easily.
Advantages: No weak section in the bar; immediate
support after installation (when resin is used); high
load and displacement capacities. Reasonable energy
absorption capacity - for a 1 m section of debonded Faceplate
#6 threadbar, 90 mm of displacement is theoretically
Nut
possible and the static energy the bolt can absorb is
around 13 kJ.
Disadvantages: The thread may not have sufficient
strength to maximize the energy absorption potential
of the bolt. Test results by Player et al. (2009a) Technical data:
showed that for toe anchored 20 mm threadbars Steel designation: Grad 75
(which means that the collar load was applied directly Yield stress: 500 MPa
onto the domed plate), the absorbed energy was from
Steel diameter: 19 mm (#6), 22 mm (#7), and 25
0.8 to 1.1 kJ only, far less than the 13.6 to 21.8 kJ
mm (#8)
energy obtained from the tests using the double-
Yield load (min): 146 kN, 200 kN, and 263 kN
embedment split loading method. The nut and thread
were stripped and the total displacement was only 8 Length: variable
to 15 mm. Steel elongation: 12%.
Major support function: Reinforce and hold. For the Displacement at peak load: 25–45 mm (debonded)
holding function to work in its full potential in burst- Total displacement capacity: 82–106 mm
prone grounds, debonding of the bar is needed. A 2.4 (debonded)
m long threadbar installed with a 50 cm of resin-
Energy absorption capacity: 13.6–21.8 kJ (19 mm,
bonded length and a 1.7 m of free stretching length
fully debonded with a debonding length of 160
would allow a minimum of 153 mm of displacement
cm), obtained from double-embedment split tube
and hence is able to perform well in slight to
test.
moderate rockburst conditions.
Key to success: Proper resin mixing; use the right
sized hole, hole length, and resin cartridge; proper Source: Player et al. (2009a), DSI, Williams Form
debonding if the bolt is used for dynamic support. A Engineering.
minimum of 700 mm bond length is required using
resin grout.

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 281

E.26 Weld mesh

Background: Weld mesh is used for surface support,


to transfer load to rockbolts, and to keep small pieces
of loose rock and broken rock from falling. Weld
meshes used in underground operations are often in
sheets of varying widths and lengths. Galvanized
meshes are available for use in corrosive grounds.
The mesh is usually specified by wire diameter
(gauge) and wire spacing. The Imperial Wire Gauge
or the Standard Wire Gauge (S.W.G.) for some
commonly used sizes is shown below. In Canadian
mines, the widely used weld meshes are #6 and #4
gauge on a typical 100 mm square grid pattern. The
standard weld mesh used in Western Australian
mines is 5.6 mm in wire diameter on a 100 mm
square grid pattern.
Gauge number Wire diameter Wire diameter

2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
(S.W.G) (mm) (Inch)

0 8.23 0.324

4 5.893 0.232 Technical data:


6 4.877 0.192 Steel designation: Carbon steel
Yield strength: 450 MPa
9 3.658 0.144
Steel wire diameter: variable (#4 and #6 are most
commonly used)
Advantages: Inexpensive, quick to install, and easy Sheet size: variable (48" to 96" wide, 48" to 24'
to be repaired; provide immediate support after long)
installation; suitable for shotcrete applications. Load capacity: 45 kN (#4 gauge, 1.5 m  3 m sheet,
Disadvantages: Not enough load and deformation pull test); 30 kN (#6 gauge, 1.5 m  1.5 m sheet)
capacity for light weight weld meshes; installation is
Displacement at peak load: 150 mm (#4) at a 1.2
not easily automated; easily damaged by fly rock
m diamond pattern, 120 mm (#6)) at a 1.2 m
from nearby blasts.
diamond pattern.
Major support function: Retain. In addition to the
Total displacement capacity: 250 mm
retaining function, meshes transfer loads to the bolts.
Energy absorption capacity: < 10 kJ/m2 (#6
Key to success: Proper installation; sufficient mesh
gauge, drop test).
overlap. In rockburst grounds, heavy gauge mesh
(e.g., #4) should be used to increase the surface
support capacity. Heavy gauge meshes are not only Source: Kaiser et al. (1996), Ortlepp and Stacey
stronger, but also stiffer which is critical to transfer (1998), DSI.
dynamic load effectively to yielding bolts. Avoid
sharp plate edges damaging the mesh wires. Using
yielding bolts can avoid damage to the mesh and thus
increase the overall capacity of the support system.

 
Draft manuscript – Copyright protected – Cai and Kaiser 2018
282 References

E.27 Yield‐Loc bolt

Background: Yield-Loc bolt is a new yielding bolt


designed for bursting and squeezing ground support.
The bolt diameter is 17.2 mm, and the steel bar is
2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

upset at one end and partially or fully encapsulated in


a polymer coating and the other end is threaded for
tensioning with a nut. The angled polymer coating Polymer coating
serves the function of resin mixing. As the bolt is
loaded, the upset head can plow through the polymer
coating and the grout, thus provide load resistance
and absorb dynamic energy.
Advantages: Improved consistency as the upset is
plowing through the polymer coating and the grout;
performance is less dependent on resin; immediate
support after installation.
Plate & nut
Disadvantages: Need to pre-define polymer coating
length; steel stretching increases with impact energy;
high cost; potential deterioration of polymer coating
if the bolt is stored for a long period.
Major support function: Dynamic energy absorption
Technical data:
and hold.
Steel designation: Grade 75
Key to success: Proper installation is important to
secure the polymer coating anchorage; need case Yield stress (typical): 525 MPa
histories to demonstrate the bolt’s performance in Steel diameter: 17.2 mm
large rockbursts. Yield load (min): 110 kN
Length: variable
Steel elongation: 10%
Displacement at peak load: 40–50 mm
Total displacement capacity: > 200 mm (depends
on the polymer coating length)
Energy absorption capacity: 16.4 kJ for about 200
mm of displacement (impact at the plate).

Source: Jennmar, Wu et al. (2010).

Draft manuscript – Copyright protected – Cai and Kaiser 2018


Rockburst Support Reference Book (I) 283

References

Bucher, R., Roth, A., Roduner, A., and Temino, J. 2010. Ground support in high stress mining with high-tensile
chain-link mesh with high static and dynamic load capacity. In 5th International Seminar on Deep and High
Stress Mining. Edited by M. Van Sint Jan and Y. Potvin, Santiago, Chile. pp. 273-282.
Cai, M., and Champaigne, D. 2009. The art of rock support in burst-prone ground. In RaSiM 7: Controlling
Seismic Hazard and Sustainable Development of Deep Mines. Edited by C.A. Tang. Rinton Press. pp. 33-46.
Cai, M., and Champaigne, D. 2012. Influence of bolt-grout bonding on MCB conebolt performance. Int. J. Rock
Mech. Min. Sci. 49(1): 165-175.
Cai, M., Champaigne, D., and Kaiser, P.K. 2010. Development of a fully debonded conebolt for rockburst support.
In 5th International Seminar on Deep and High Stress Mining. Edited by M. Van Sint Jan and Y. Potvin,
Santiago, Chile. pp. 329-342.
Doucet, C., and Gradnik, R. 2010. Recent developments with the RoofexTM bolt. In 5th International Seminar on
Deep and High Stress Mining. Edited by M. Van Sint Jan and Y. Potvin, Santiago, Chile. pp. 353-366.
Falmagne, V., and Simser, B.P. 2004. Performance of rockburst support systems in Canadian mines. In Ground
Support in Mining and Underground Construction. Edited by E. Villasescusa and Y. Potvin. pp. 313-318.
He, M., Gong, W., Wang, J., Qi, P., Tao, Z., Du, S., and Peng, Y. 2014. Development of a novel energy-absorbing
bolt with extraordinarily large elongation and constant resistance. Int. J. Rock Mech. Min. Sci. 67: 29-42.
He, M., Xia, H., Jia, X., Gong, W., Zhao, F., and Liang, K. 2012. Studies on classification, criteria and control of
rockbursts. Journal of Rock Mechanics and Geotechnical Engineering 4(2): 97-192.

2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo
Hoek, E., and Wood, D. 1987. Support in underground hard rock mines. In Underground Support Systems. Edited
by J. Udd. Canadian Institute of Mining and Metallurgy. pp. 1-6.
Hutchinson, D.J., and Diederichs, M.S. 1996. Cablebolting in Underground Mines. BiTech Publishers Ltd.
Jager, A.J. 1992. Two new support units for the control of rockburst damage. In Rock Support in Mining and
Underground Construction. Edited by P.K. Kaiser and D.R. McCreath. pp. 621-631.
Kaiser, P.K., Tannant, D.D., and McCreath, D.R. 1996. Canadian Rockburst Support Handbook. Geomechanics
Research Centre, Laurentian University, Sudbury, Ontario.
Kirsten, H.A.D., and Labrum, P.R. 1990. The equivalence of fibre and mesh reinforcement in the shotcrete used in
tunnel-support systems. J. South Afr. Inst. Min. Metall. 90(7): 153-171.
Labrie, D., Doucet, C., and Plouffe, M. 2008. Design guidelines for the dynamic behaviour of ground support
tendons.
Li, C. 2010. A new energy-absorbing bolt for rock support in high stress rock masses. Int. J. Rock Mech. Min. Sci.
47: 396-404.
Li, C., and Charette, F. 2010. Dynamic performance of the D-Bolt. In Proc. 5th Int. Seminar on Deep and High
Stress Mining. Edited by M. Van Sint Jan and Y. Potvin. pp. 321-328.
Li, T., Brown, E.T., Coxon, J., and Singh, U. 2004. Dynamic capable ground support development and application.
In Ground Support in Mining and Underground Construction. Edited by E. Villasescusa and Y. Potvin. Taylor
& Francis Group, London. pp. 281-288.
Li, T., Brown, E.T., Singh, U., and Coxon, J. 2003. Dynamic support rationale and systems. In ISRM 2003
Technology roadmap for rock mechanics. South African Institute of Mining and Metallurgy. pp. 763-768.
Morton, E.C., G., T.A., and Villaescusa, E. 2009. The performance of mesh, shotcrete and membranes for surface
ground support. In ROCKENG09: Proceedings of the 3rd CANADA-US Rock Mechanics Symposium. Paper
4002.
Oler, R. 2012. DSI new developments in yieldable rock bolts. In Dynamic Ground Support Application
Symposium.
Ortlepp, W.D. 1994. Grouted rock-studs as rockburst support: A simple design approach and an effective test
procedure. J. South Afr. Inst. Min. Metall. 94: 47-53.
Ortlepp, W.D. (ed). 1997. Rock Fracture and Rockbursts – An Illustrative Study. The South African Institute of
Mining and Metallurgy, Johannesburg.
Ortlepp, W.D., Bornman, J.J., and Erasmus, P.N. 2001. The Durabar - a yieldable support tendon - design rationale
and laboratory results. In 5th lnt. Symp. on Rockburst and Seismicity in Mines. pp. 263-266.
Ortlepp, W.D., and Erasmus, P.N. 2005. Dynamic testing of a yielding cable anchor. In 3RD Southern African Rock
Engineering Symposium.
Ortlepp, W.D., and Stacey, T.R. 1994. The need for yielding support in rockburst conditions, and realistic testing of
rockbolts. In Proceedings International Workshop on Applied Rockburst Research, Santiago, Chile.
Draft manuscript – Copyright protected – Cai and Kaiser 2018
284 References
Ortlepp, W.D., and Stacey, T.R. 1998. Performance of tunnel support under large deformation static and dynamic
loading. Tunnelling and Underground Space Technology 13(1): 15-21.
Ortlepp, W.D., and Stacey, T.R. 1999. Retainment support for dynamic events in mines. In Rock Support and
Reinforcement Practice in Mining. Edited by E. Villaescusa and C.R. Windsor and A.G. Thompson. A.A.
Balkema. pp. 329-333.
Ortlepp, W.D., and Swart, A.H. 2002. Performance of various types of containment support under quasi-static and
dynamic loading conditions, Part II.
2018 Unpublished limited-distribution manuscript for soliciting feedback - For personal use by Garcia Hugo

Player, J.R. 2004. Field performance of cone bolts at Big Bell mine. In Ground Support in Mining and
Underground Construction. Edited by E. Villasescusa and Y. Potvin. pp. 289-298.
Player, J.R., Thompson, A.G., and Villasescusa, E. 2008. Improvements to reinforcement systems through dynamic
testing. In Proceedings Tenth Underground Operators’ Conference. The Australasian Institute of Mining and
Metallurgy, Melbourne. pp. 79-88.
Player, J.R., Thompson, A.G., and Villasescusa, E. 2009a. Dynamic testing of threadbar used for rock
reinforcement. In ROCKENG09: Proceedings of the 3rd CANADA-US Rock Mechanics Symposium. Edited
by M. Diederichs and G. Grasselli, Toronto. Paper 4030.
Player, J.R., Villasescusa, E., and Thompson, A.G. 2009b. Dynamic testing of friction rock stabilisers. In
ROCKENG09: Proceedings of the 3rd CANADA-US Rock Mechanics Symposium. Edited by M. Diederichs
and G. Grasselli, Toronto.
Plouffe, M., Anderson, T., and Judge, K. 2008. Rock bolts testing under dynamic conditions at CANMET-MMSL.
In 6th International Symposium on Ground Support in mining and Civil Engineering Construction. pp. 581-
596.
Potvin, Y. 2009. Surface support in extreme ground conditions HEA Mesh. In SRDN 2009. Edited by P. Dight,
Perth, Australia. pp. 111-110.
Potvin, Y., and Heal, D. 2010. Dynamic testing of High Energy Absorption (HEA) mesh. In 5th International
Seminar on Deep and High Stress Mining. Edited by M. Van Sint Jan and Y. Potvin, Santiago, Chile. pp. 283-
300.
Punkkinen, A.R., and Mamidi, N.R. 2010. Effective ground support system design to manage seismic hazard in a
high stress diminishing pillar at a Vale mine. In 5th International Seminar on Deep and High Stress Mining.
Edited by M. Van Sint Jan and Y. Potvin, Santiago, Chile. pp. 367-381.
Roth, A., Windsor, C., Coxon, J., and deVries, R. 2004. Performance assessment of high-tensile steel wire mesh for
ground support under seismic condition. In Ground Support in Mining and Underground Construction. Edited
by E. Villasescusa and Y. Potvin. pp. 1107-1120.
Scott, J.J. 1977. Friction rock stabilizers - a new rock reinforcement method. In Proc. 1977 SMEAIME Annual
Meeting, Atlanta, Georgia.
Simser, B. 2007. The weakest link: Ground support observations at some Canadian Shield hard rock mines. In 4th
International Seminar on Deep and High Stress Mining, Perth.
Stacey, T.R., and Ortlepp, W.D. 2001. Tunnel surface support capacities of various types of wire mesh and
shotcrete under dynamic loading. J. South Afr. Inst. Min. Metall.: 337-342.
Tannant, D.D. 1997. Shotcrete performance near blasts.
Varden, R., Lachenicht, R., Player, J., Thompson, A., and Villaescusa, E. 2008. Development and implementation
of the Garford Dynamic Bolt at the Kanowna Belle Mine. In 10th Underground Operators’ Conference. pp. 95-
102.
Villaescusa, E., and Wright, J. 1997. Permanent excavation reinforcement using cement grouted split set bolts. In
The AusIMM Proceedings 1997. 1 pp. 65-69.
Wu, Y., Oldsen, J., and Lamothe, M. 2010. The Yield-Lok bolt for bursting and squeezing ground support. In 5th
International Seminar on Deep and High Stress Mining. Edited by M. Van Sint Jan and Y. Potvin, Santiago,
Chile. pp. 301-308.

Draft manuscript – Copyright protected – Cai and Kaiser 2018

You might also like