UNSW MINE3310 Rock Mechanics

Download as pdf or txt
Download as pdf or txt
You are on page 1of 215

MINING GEOMECHANICS

MINE3310: MINING GEOMECHANICS

The University of New South Wales

Rock Mechanics – Course Reader

OFFICAL COURSE READER: FOR USE IN ROCK MECHANICS EXAM.

ANNOTATIONS TO THIS DOCUMENT ARE PERMISSABLE FOR THE EXAM.

Version 1.10:Jan 2010: UNSW: GS 1-1


Owner List
The Owners-SP 2657

Lot Unit Name and Main Contact Address Levy Address Notice Address Entitlement Balance
(-)prepaid
1 1 SAMIR BAHIG SHOUSHA 1 -1.46
2/62 MASCOT DRIVE, EASTLAKES NSW 2018 2/62 MASCOT DRIVE, EASTLAKES NSW 2018 2/62 MASCOT DRIVE, EASTLAKES NSW 2018
2 2 MR S B SHOUSHA 1 -2.45
2/62 MASCOT DRIVE, EASTLAKES NSW 2018 2/62 MASCOT DRIVE, EASTLAKES NSW 2018 2/62 MASCOT DRIVE, EASTLAKES NSW 2018
3 3 EDWARD ADHIKARI & DAISY BISWAS 1 0.00
3/62 MASCOT DRIVE, EASTLAKES NSW 2018 3/62 MASCOT DRIVE, EASTLAKES NSW 2018 3/62 MASCOT DRIVE, EASTLAKES NSW 2018
4 4 SEAN & CATHERINE O'DONOGHOE 1 0.00
4/62 MASCOT DRIVE, EASTLAKES NSW 2018 4/62 MASCOT DRIVE, EASTLAKES NSW 2018 4/62 MASCOT DRIVE, EASTLAKES NSW 2018
5 5 MR DAVID HENRY POPPLEWELL 1 0.00
5/62 MASCOT DRIVE, EASTLAKES NSW 2018 5/62 MASCOT DRIVE, EASTLAKES NSW 2018 5/62 MASCOT DRIVE, EASTLAKES NSW 2018
6 6 MR G & MRS U KASSAR 1 0.00
2 KYOGLE STREET, EASTLAKES NSW 2018 2 KYOGLE STREET, EASTLAKES NSW 2018 2 KYOGLE STREET, EASTLAKES NSW 2018
7 7 MR J & MRS M REPETTO 1 0.00
42 HIGHCLERE AVENUE, ROCKDALE NSW 42 HIGHCLERE AVENUE, ROCKDALE NSW 42 HIGHCLERE AVENUE, ROCKDALE NSW
2216 2216 2216
8 8 TRAM THI NGOG LE 1 0.00
8/62 MASCOT DRIVE, EASTLAKES NSW 2018 8/62 MASCOT DRIVE, EASTLAKES NSW 2018 8/62 MASCOT DRIVE, EASTLAKES NSW 2018
9 9 MR & MRS MAHMOUD 1 404.80
9/62 MASCOT DRIVE, EASTLAKES NSW 2018 9/62 MASCOT DRIVE, EASTLAKES NSW 2018 9/62 MASCOT DRIVE, EASTLAKES NSW 2018
10 10 YOHANES SOE & HANG LIN WIE 1 404.80
10/62 MASCOT DRIVE, EASTLAKES NSW 10/62 MASCOT DRIVE, EASTLAKES NSW 10/62 MASCOT DRIVE, EASTLAKES NSW
2018 2018 2018
11 11 MESSRS T, A & J & MRS K JAKUS 1 0.00
11/62 MASCOT DRIVE, EASTLAKES NSW 11/62 MASCOT DRIVE, EASTLAKES NSW 11/62 MASCOT DRIVE, EASTLAKES NSW
2018 2018 2018
12 12 MRS Y KASSAR 1 0.00
68 FLORENCE AVENUE, EASTLAKES NSW 68 FLORENCE AVENUE, EASTLAKES NSW 68 FLORENCE AVENUE, EASTLAKES NSW
2018 2018 2018

28/08/2009 18:49 Mark Gitman MG Strata and BMC Management Pty Ltd Page 1
Rock
MINING GEOMECHANICS

Mechanics

1
TOPIC 1:
INTRODUCTION TO
ROCK MECHANICS
TOPICS

1.1 Introduction to Mining Rock Mechanics


Rock mechanics can be described as being;

“the theoretical and applied science of the mechanical


behaviour of rock and rock masses; it is that branch of
mechanics concerned with the response of rock and
rock masses to the force fields of their physical
environment.”

following US National Committee of Rock Mechanics, in Brady & Brown,


1999. Rock mechanics as a discipline forms part of the broader subject
of geomechanics, which is concerned with the mechanical responses of
all geological materials, including soils.

Engineering rock mechanics when applied to mining engineering (an


understanding of which is the main purpose of this course) is;

“concerned with the application of the principles of


engineering mechanics to the design… of excavations
in mines”

following Brady & Brown, 1999. Some examples of such excavations


are open pit mine slopes, vertical mine shafts, horizontal mine
development excavations and production stopes.

In summary, the science of rock mechanics is the science of the


mechanical behaviour of rock and rock masses. It impacts on mine

Version 1.10:Jan 2010: UNSW: GS 1-2


MINING GEOMECHANICS

safety, productivity and resource recovery, and thus is fundamental to


both the design and operating phases of a mine’s life.

1.1.1 Learning objectives


By the end of this introductory topic you will:

 Be able to appreciate the importance of rock mechanics in all


facets of a mining operation; and
 Be able to list the key issues and questions when considering
any mining rock mechanics problem.

1.2 Design in a Rock -Mass or –Material?


Consider the figure below. The rock mass is the combination of the
intact rock material and the rock defects or discontinuities. This is the
engineering material that we work with and it’s behaviour will influence
the engineering designs we develop.

The intact rock material is the solid part of the rock mass between the
geological structures (rock defects or discontinuities). The ‘intact’ parts
gives rise to the alternate name of ‘intact rock’.

In rock mechanics, the term discontinuity refers to all joints, faults, and
planes of weakness that bound blocks of intact rock material. A term
that is used in some texts is ‘fractures’.

Intact

Design and analysis of excavations in rock require an understanding of


the behaviour of both the intact rock material AND the discontinuities in
the rock mass as a whole. When considering the rock mechanics of

Version 1.10:Jan 2010: UNSW: GS 1-3


MINING GEOMECHANICS

any mining engineering problem, the following issues need to be


addressed:

 Excavation Size (& Shape)


 Rock (mass) Structure
 Rock (mass) Strength
 Rock (mass) Stress (& Strain).

For the purposes of memory, we will abbreviate these to ‘The 4 ‘S’s’.


That is (in order):

Size (and/or Shape);


Structure;
Strength; and
Stress (or Strain).

Size and shape are sometimes collectively referred to as geometry.


Strictly speaking, geometry should be used in reference to shape and
size should refer scale of the excavation.

1.2.1 Size of excavation


Size should be considered in relative and comparative terms. For
example: the size (length and width) of a joint or shear zone compared
to the size of a drive, the size of a fault or bedding plane compared to
the size of a stope or open pit batter. What you need to consider is the
volume of interest. It is best demonstrated using the figure on the next
page. Consider the circles in the figure. If 1 represents the smallest
circle (both underground and in the pit) and 5 represents the largest
circle, then, consider how each of the scales shown in the table below
might impact on;

A. An underground drive;
B. An open pit bench;
C. A large underground stope 60m high by 20m by 20m; and
D. A pit slope 60m high containing 10 benches.

Version 1.10:Jan 2010: UNSW: GS 1-4


MINING GEOMECHANICS

Effect of size (scale) relative to the excavation (after Hoek, 1983)

Circle Underground Open Pit


size
1 A cylinder of rock the size A cylinder of rock the size of
of a soft drink can a load of bread
2 A cylinder of rock the size A volume the size of a
of a loaf of bread coffee table containing a
containing one joint single joint
3 A volume the size of a A volume with diameter
coffee table equal to the height of one
bench
4 A volume with a diameter A volume with diameter
equal to the height of the equal to the height of two
drive containing 10 joint benches containing 10 joint
sets sets
5 A volume with a diameter A volume with diameter
equal to the height of 2 equal to the height of six
drives containing >10 benches containing >10
joint sets joint sets

Version 1.10:Jan 2010: UNSW: GS 1-5


MINING GEOMECHANICS

We have already learnt that size and shape are generally considered
collectively. We have also learnt that, strictly speaking, shape is defined
by geometry. We will therefore use this and consider geometry in three-
dimensional space; namely length, breadth, depth and orientation.

Shape has a major impact on stability, both in underground and surface


excavations. For example in sub level open stoping, stability is achieved
by having high vertical and short horizontal dimensions, or low horizontal
and short vertical dimensions, as shown below.

Stable shapes for sublevel stoping (after Villaescusa, 2000)

1.2.2 Structure
For the purposes of the module, we will use this term to refer to the
major and minor discontinuities (fractures, etc.) that make up the rock
mass.

Again we are concerned with the geometry of the geological structures


in three-dimensional space: length, width (or height), aperture (or infill
thickness) and orientation. For example the orientation of a fault will
effect the direction of the principal stresses. We should also consider
the orientation in a relative sense. For example, a joint set that is
dipping out of the face creates less favourable conditions for horizontal
mining compared to a joint set dipping into the face (as illustrated in the
next figure).

Version 1.10:Jan 2010: UNSW: GS 1-6


MINING GEOMECHANICS

Another example would be the direction and orientation of structures


coinciding with an open pit slope face. Are they dipping out (and
creating a possible sliding failure?) or dipping in (possible toppling
failure?).

Joint
systems or
bedding
dipping out of
face

Diagram representing the effect of a structure dipping out of a slope face


– potential for sliding to occur on the structure or discontinuity.

Joint
systems or
bedding
dipping
into face

Diagram representing the effect of structures dipping into face –


potential for toppling failure (cf dominoes in a row)

In combination with this question, we are also interested in the


frequency (the number of times they occur per unit length or core or unit

Version 1.10:Jan 2010: UNSW: GS 1-7


MINING GEOMECHANICS

volume of rock mass), persistence (how far they extend without a break;
that is, their length), and their shear strength.

1.2.3 Strength
The behaviour of slopes and underground spans is largely controlled by
the strength of the rock mass, which in turn depends upon the
geometrical nature and strength of the geological discontinuities and
intact rock material.

In this instance we are referring to the overall rock mass strength rather
than that of any one particular structure. We must question whether the
rock mass can sustain the design we are trying to implement; for how
long and what support it needs for the life of the structure.

However, where specific discontinuities control stability (for instance, the


joint system or bedding dipping out of the slope face illustrated
previously), then the discontinuity strength itself is a primary
consideration.

In both cases though, in the analysis of excavation stability, we need to


understand, not only the strength or the rock mass but also the stress
being placed on it …. Clearly, when the applied stress exceeds the
available strength of the rock mass, then failure will initiate. Generally,
for mining excavations, failure cannot be tolerated.

1.2.4 Stress
In simple terms, stress refers to the forces that are applied to the rock
mass. The magnitude of the stress is related to a number of complex
interactions, including the depth in the rock mass below ground surface
we are concerned with (similar to water pressure) and tectonic activity
and other aspects of the geological history that the rock mass has
experienced to date.

When we create a mine, the excavations disturb the in situ stress field,
and the new stress field that is created is termed the induced stress
field. An analogue to induced stress around an excavation is the flow of
river water around a bridge post. The stresses build up (concentrate on
the sides of the excavation) parallel to the direction of the maximum
stress direction.

The specific induced stress field is both a function of the parameters


named above, and also excavation shape and location relative to other
underground excavations.

Version 1.10:Jan 2010: UNSW: GS 1-8


MINING GEOMECHANICS

Induced stress flow around a drive (profile cross-section)

Stress is a major issue in underground mines, especially deep level


mines. We must consider if the strength of the rock mass (see above) is
enough to sustain the stresses placed on it. Excavations will often
possess;

 zones of over-stress where the stress may be close to or greater


than the compressive strength of rock mass. High stresses may
cause rupture of the rock mass.
 zones of under-stress (relaxation) where the stress may be close
to or greater than the tensile strength of rock mass. Low stresses
allow joints to dilate (move apart) and blocks to fall or unravel
under the impact of gravity.

Strain refers to the relative displacement (to the original position) of


particles within a rock mass due to changes in the stress state and can
be either recoverable or permanent.

Changes in the stress state can occur due to changes in temperature


conditions, tectonic conditions and groundwater conditions, or due to
dynamic loads such as seismicity or blasting. The relationship between
stress and strain allows us to measure strain and determine the stress
at that position in the ground.

1.3 Rock Mechanics and Underground Mining


Engineering
Irrespective of the underground mining technique used (surface
excavation design is covered in Geotechnics 431), there are four
common rock mechanics objectives for the performance of a mine
structure, and three different types of mine excavations. The types of
excavations are generally classed as;

Version 1.10:Jan 2010: UNSW: GS 1-9


MINING GEOMECHANICS

Mine access and service excavation


The main shaft, level drives and cross cuts, ore haulages, ventilation
shafts and airways. Their duty life is comparable with, or exceeds, the
mining life of the orebody and they are usually developed in barren
ground.

Service and operating excavations


Directly associated with ore recovery and consist of the access cross
cuts, drill headings, access raises, extraction headings and ore passes,
from, or in which, various ore production operations are undertaken.
These openings are developed in the orebody, or in country rock close
to the orebody boundary, and their duty life is limited to the duration of
mining activity in their immediate vicinity. Many such excavations are
eliminated by the mining operation.

Ore source
It may be a stope, with well-defined, free-standing rockwalls forming the
geometric limits for the mined void, which increases in size with the
progress of ore extraction. Alternatively, the ore source may be a rubble-
filled space with fairly well-defined lower and lateral limits, usually
coincident with the orebody boundaries. The rubble is generated by
inducing disintegration of the rock above the crown of the orebody,
which fills the mined space as extraction proceeds. The lifetime of these
different types of ore source openings is defined by the duration of
active ore extraction.

The performance objectives are specifically;

 to ensure the overall stability of the complete mine structure,


defined by the main ore sources and mined voids, ore remnants
and adjacent country rock.
 to protect the major service openings throughout their designed
duty life
 to provide secure access to safe working places in and around
the centres of ore production;
 to preserve the mineable condition of unmined ore reserves.

To achieve these objectives, rock mechanics activities need to be


conducted within an organisational framework that permits the
exchange and integration of concepts, requirements, information and
advice from and between management, geologists, planning engineers,
production personnel and rock mechanics engineers, as shown below.

Version 1.10:Jan 2010: UNSW: GS 1-10


MINING GEOMECHANICS

Once this framework is in place, the general rock engineering process


can begin. The figure below illustrates the process and also forms the
basis for the following topics in this Module.

Site Characterisation (Data collection and Analysis)

Mine model formulation (Engineering Analysis)

Design (Global & local)

Implementation

Rock Mass performance monitoring

Retrospective Analysis

Version 1.10:Jan 2010: UNSW: GS 1-11


MINING GEOMECHANICS

References
Whilst there is no designated text book for this Module, the following
texts have been used extensively to develop this Module material and
are strongly recommended both for the purposes of this course and for
your future professional career. Throughout each Topic, additional
references have been given, but for clarity and readability, references to
the following source documents have been omitted.
rd
Brady, B H G and Brown, E T (2004). Rock Mechanics for Underground Mining. 3
Edition, Springer, Netherlands.
Goodman, R E (1980). Introduction to Rock Mechanics, John Wiley and Sons,
Toronto, Canada.
Hudson, J A and Harrison, J P (1997). Engineering Rock Mechanics, An
Introduction to the Principles, Pergamon, Elsevier Science, UK.
Hoek, E and Brown, E T (1997). Underground Excavations in Hard Rock,
Chapman and Hall, UK.
Hoek, E and Bray, J W (1997). Rock Slope Engineering, Chapman and Hall,
London.
Weijermars, R (1997). Principles of Rock Mechanics, Alboran Science Publishing,
Amsterdam.

Notes from the following source have been used on occasion:

Hoek, E. Rock Engineering. Course Notes. Available via www.rocscience.com

The following texts are also considered useful for specific areas of the
course and may give you an alternative perspective to the course
material:
Bieniawski, Z T (1984). Rock Mechanics Design in Mining and Tunnelling.
Balkema, Rotterdam.
Bieniawski, Z T (1989). Engineering Rock Mass Classifications. Wiley, New York.
Hoek, E, Kaiser, P K and Bawden, W F (1995). Support of Underground
Excavations in Hard Rock. Balkema, Rotterdam.
Kirkaldie, L (Ed) (1988). Rock Classification for Engineering Purposes. American
Society for Testing and Materials (ASTM), Philadelphia.
Lowrie, R L (Ed) (2002). SME Mining Reference Handbook, Society for Mining,
Metallurgy, and Exploration, Littleton, Co., USA
Singh, B and Goel, R K (1999). Rock Mass Classification- A Practical Approach
in Civil Engineering. Elsevier, Oxford, UK.

Version 1.10:Jan 2010: UNSW: GS 1-12


MINING GEOMECHANICS

Acknowledgements
The course is supported by industry through the Minerals Council of
Australia (MCA), in collaboration with Mining Education Australia (MEA).
MEA is a four-way joint venture between Curtin University of
Technology, The University of Adelaide, The University of New South
Wales and The University of Queensland.

The original (2007) Reader was produced through the MEA and the
following authors’ contributions are acknowledged;

Dr Habib Alehossein
Professor Bruce Hebblewhite
Dr Shivakumar Karekal
Professor Peter Lilly,
Dr Hamid Nikraz
Dr Glenn Sharrock
Mr Simon Thomas
Dr Alan Thompson
Professor Ernesto Villaescusa
Dr John Watson
Mr Chris Windsor

Revision 2010
The revised Reader (2010) draws heavily from the 2007 document and
contains additional contributions from;

Professor Stephen Priest, The University of Adelaide.


Dr Glenn Sharrock, The University of New South Wales.
Professor Roger Thompson, Curtin University of Technology, WASM

Version 1.10:Jan 2010: UNSW: GS 1-13


Rock
MINING GEOMECHANICS

Mechanics

2
TOPIC 2: ROCK MASS
CLASSIFICATION
METHODS

2.1 Introduction: Classification vs.


Characterisation
It is important to define the terms „classification‟ and „characterisation‟
because they have different applications in rock mechanics.

Classification is defined as;

‘the action or process of classifying something


according to shared qualities or characteristics’.

Thus, classification is the result of putting objects into different classes.


The purpose of such classification is get a better overview of a
phenomenon or set of data, and to try to gain an improved
understanding of them.

By contrast, characterisation is defined as;

‘describing the distinctive nature or features of


something’

such as intact rock characteristics, discontinuity characteristics and the


density and pattern of discontinuities. Characterising the rock mass and
defining its behaviour is the key step to safe and economic design and
extraction in any underground or open pit operation. The rock mass to
be considered is often initially divided into geological or preferably
geomechanical domains or zones (e.g. hangingwall granite, hangingwall
schist, ore body, footwall gabbro, main fault, etc) and each domain
should be classified separately. If during the classification process

2-1
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

significant changes in structural character are noted, then additional


subdivisions should be created until all geomechanical domains are
identified and assigned properties.

In practical rock mechanics, the task is to:

 Identify the features or parameters of importance or relevance to


a project and the assessments to be performed; and
 Measure and/or describe the properties of these parameters,
giving them values or ratings according to their structure,
composition and properties.

From these definitions you will note that the classification process is
driven by two questions:

1. What is the purpose of the classification?


2. What method do I use commensurate with the purpose?

The diagram below shows how characterization leads to the application


of classification systems – but only as one application option of several.

Classification systems have been designed on case histories and are


always appropriate for their original application, however considerable
caution must be exercised in applying rock mass classifications to other
rock engineering problems.

The general rock engineering process diagram (introduced in Topic 1)


shows that Classification Schemes would fall into the first three
categories. The broader topic of rock mass characterisation and

2-2
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

behaviour spans the first two levels of the general rock engineering
process. Rock classification is part of the second step of the general
process and the result generated from a classification analysis can be
used in the third process, that of initial design (BUT ideally only as an
adjunct, not a complete design).

Site Characterisation (Data collection and Analysis)

Mine model formulation (Engineering Analysis)

Design (Global & local)

Implementation

Rock Mass performance monitoring

Retrospective Analysis

2.1.1 Learning Objectives


By the time you have completed this topic you will be able to:

1. Differentiate between classification and characterisation;


2. Compare and contrast the RQD, RMR, Q and GSI system
classification schemes;
3. Apply the RQD, RMR, Q and GSI systems classification
schemes to classify a rock mass;
4. Calculate excavation stand-up times, support requirements,
rock mass modulus and GSI from classification data.

2-3
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

2.2 Introduction to Classification Methods


Most Australian underground mines use classification systems to
characterise the ground conditions for geotechnical design. Rock mass
classification schemes can be beneficial at the feasibility and preliminary
design stages of a mine, when very little detailed information on the rock
mass and its stress and hydrologic characteristics is available.
Classification can be applied at two levels;

 At the simplest level, this may involve using classification


systems as a check-list to ensure that all relevant information has
been considered.
 At a more advanced level, one or more rock mass classification
schemes may be used to build up a picture of the composition
and characteristics of a rock mass to provide initial estimates of
support requirements; and/or estimates of the strength and
deformation properties.

It is important to understand that the use of rock mass classification


systems cannot replace engineering analysis and design procedures.
However, engineering design procedures require access to relatively
detailed information, none of which may be available at an early stage in
the project. Therefore, classification can be used as a tool to assess the
"quality" of rock masses. In the absence of other data, these tools can
be used to establish rock mass parameters, or guide the selection of
reinforcement and support methods.

Different classification systems place different emphases on the various


parameters, and it is recommended that at least two methods be used
at any site during the early stages of a project.

The two best-known systems are Barton‟s Rock Tunnelling Quality


Index (Q-System, also known as the NGI System) and Bieniawski‟s
Rock Mass Rating (RMR, also known as the Geomechanics
Classification). The RMR makes no allowance for high ground stresses
at depth, and thus as underground mines trend deeper, Barton‟s Q-
System is more likely to be used. The GSI (Geological Strength Index)
(Hoek, 1994) is also introduced in which the strength of a discontinuous
or jointed rock mass is analysed.

Common to both the RMR and NGI Q-system is the Rock Quality
Designation (RQD) and this is introduced first.

2-4
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

2.3 Rock Quality Designation Index (RQD)


The Rock Quality Designation index (RQD) was developed by Deere
et al in 1967 to provide a quantitative estimate of rock mass quality
from drill core logs. RQD is defined as the percentage of intact core
pieces longer than 100 mm in the total length of core (measured
along core centreline), as illustrated below.

35

 Length of core pieces  100 mm


RQD  x100
Total length of core run

RQD is intended to represent the rock mass quality in situ. When using
diamond drill core, care must be taken to ensure that fractures, which
have been caused by handling or the drilling process, are identified and
ignored when determining the value of RQD. Hence, in the above
figure, the section of core (L=35cm) the drilling break is ignored in
determining RQD and L=35cm terminates at a fracture.

When no core is available but discontinuity traces are visible in surface


exposures or exploration adits, the RQD may be estimated from the
number of discontinuities per unit volume (Jv). The suggested
relationship for clay-free rock masses is:

RQD  115 3.3J v

2-5
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

RQD is a directionally dependent parameter and its value may change


significantly, depending upon the borehole orientation. The use of the
volumetric joint count can be quite useful in reducing this directional
dependence. Additionally, since RQD is a measure of fracture
frequency (λ fractures/meter), there is a relationship between the two
such that;

RQD  3.68  110.4

In the context of this discussion, the most important use of RQD is as a


component of the RMR and Q rock mass classifications.

2.4 Rock Mass Rating System (RMR)


As you are reading this section, refer to Appendix A for a copy of the
RMR table. Bieniawski (1988) identified the aims of the RMR
system:

1. To identify the most significant parameters influencing the


behaviour of a rock mass;
2. To divide a particular rock mass formation into a number of rock
mass classes of varying quality;
3. To provide a basis for understanding the characteristics of each
rock mass class;
4. To derive quantitative data for engineering design; and
5. To provide a common basis for communication between
engineers and geologists.

2.4.1 RMR Parameters


The following six RMR parameters are used to classify a rock mass.
Refer to Appendix A for the RMR table and sub-sections (A1-A5, B,
E, F)

1. Uniaxial compressive strength of rock material (A1). Rating


from 0 to 15.
2. Rock quality Designation (RQD) (A2). Rating from 3 to 20.
3. Spacing of discontinuities (A3). Rating from 5 to 20.
4. Condition of discontinuities (A4 as described in section E).
Rating from 0 to 30.
E1 Discontinuity length. Rating from 0 to 6.
E2 Discontinuity spacing. Rating from 0 to 6.
E3 Discontinuity roughness. Rating from 0 to 6.

2-6
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

E4 Discontinuity infill. Rating from 0 to 6.


E5 Discontinuity weathering. Rating from 0 to 6.
5. Groundwater conditions (A5). Rating from 0 to 15.
6. Orientation of discontinuities (B as described in section F).
Rating adjustment from -50 to 0.

The rating is an outcome of a classification of each parameter. The


ratings of each of these parameters are summed to give a value of
RMR. The calculated RMR value may be used to find which of five pre-
defined rock mass classes the rock mass belongs to (going from very
good rock to very poor rock). Refer to Appendix A for the rock mass
classes (sections C, D of the Table).

The best approach is to rate parameters 1, 2, 3, 4 & 5 and determine the


UNADJUSTED RMR. Parameter 6 (B) can then be rated and used to
adjust the RMR to get the FINAL ADJUSTED RMR. The reason for this
two-stage approach is that the rating (B) for discontinuity orientation is a
function of the tunnel dip and strike relative to that of the discontinuities
and, if the tunnel dip and strike change (as they would in practice), a
different RMR will result for each change of direction.

5
RMR   Ai  B
i 1

In applying this classification system, the rock masses are divided into a
number of structural regions AND EACH REGION CLASSIFIED
SEPARATELY. The boundaries of the structural regions usually
coincide with major structural features. However, from the practical
point of view, the rating is also related to length of the blasting round or
the recently excavated tunnel section.

When mixed-quality rock masses are encountered, it is essential to


identify the „most critical‟ condition. This means that the geological
features exercising the greatest influence on stability will have the over-
riding influence. For example, a fault or a shear in a high quality rock
will play the dominant role irrespective of the high intact rock material
strength.

Where several joint sets are identified, it is the joint set that influences
tunnel stability the most that should be used in the assessment of RMR.
For example, in a tunnel, a joint set whose strike is parallel to the tunnel
axis would exert the greatest influence on stability. The summed ratings
of the classification parameters for THIS discontinuity set will constitute
the overall RMR.

2-7
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

Where no one discontinuity set is dominant or of critical importance, or


when estimating rock mass strength and deformability, the ratings from
each of the discontinuity set classifications are averaged for the
appropriate individual classification parameter.

2.4.2 Stand-up Time


Bieniawski published a set of guidelines for estimating the stand-up time
and for selecting rock support in tunnels, based on the RMR value. The
concept of stand-up time represents the duration of time an excavation
will remain serviceable; after which significant instability and caving
occurs. Within the shaded zone the RMR contour line gives the
anticipated stand-up time without support.

Bieniawski also published a set of guidelines for the selection of support


in tunnels in rock for which the value of RMR has been determined.
These guidelines are reproduced in the table on the next page. Note
that these guidelines have been published for a 10m span horseshoe
shaped tunnel, constructed using drill and blast methods, in a rock mass
subjected to a vertical stress <25 MPa (equivalent to a depth below
surface of approx. <900 m).

Where support is indicated in the table overleaf, permanent roof


support pressure (Proof) (kPa) can also be estimated from;

100 RMR
Proof  B r
100

Where B refers to the tunnel width (m) and ρr the rock density
(kg/m3).

2-8
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

The support factor of Safety (FOS) can then be defined as;

Fb
FOS  2
Proof xSb

Where Fb refers to the bolt (tension) capacity (kN) and Sb the bolt
spacing (m).

(NOTE: In the support table, wire mesh is more commonly replaced with steel
fibre reinforced shotcrete these days).

2.4.3 Strengths and limitations


The system works well to classify the rock mass quality, since it is
relatively well defined and the rating for each parameter can be
estimated with acceptable precision. The relatively small database
makes the system less applicable to be used as an empirical design
method for rock support.

The output from the RMR classification method tends to be rather


conservative, which can lead to over design of support systems.

2-9
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

2.5 Q System
As you are reading this section, refer to Appendix B for copies of the Q-
System tables.

On the basis of an evaluation of 212 case histories from 50 different rock


types, Barton (of the Norwegian Geotechnical Institute (NGI)) proposed
a Tunnelling Quality Index (Q) as a classification system for estimating
rock support in tunnels. It is a quantitative classification system based
on a numerical assessment of the rock mass quality.

The numerical value of the index Q is defined by six parameters and the
following equation:

RQD J r J
Q x x w
Jn J a SRF

where:

RQD is the rock quality designation (RQD as a value of 1 to100);

Jn is the joint set number;

Jr is the joint roughness number;

Ja is the joint alteration number;

Jw is the joint water reduction factor; and

SRF is the stress reduction factor.

The rock mass quality can range from Q = 0.001 to Q = 1000 on a


logarithmic quality scale.

 The first quotient (RQD/Jn) represents roughly the block size of


the rock mass, with the two extreme values (100/0.5 and 10/20)
differing by a factor of 400. If the quotient is interpreted in units of
centimetres, the extreme 'particle sizes' of 200 to 0.5 cm are
seen to be crude but fairly realistic approximations. (Clay
particles are of course excluded).
 The second quotient (Jr/Ja) describes the roughness and
frictional characteristics of the rock mass joint walls or filling
materials. This quotient is weighted in favour of rough, unaltered
joints in direct contact. It is to be expected that such surfaces will

2-10
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

be close to peak strength, that they will dilate strongly when


sheared, and they will therefore be especially favourable to
tunnel stability. When rock joints have thin clay mineral coatings
and fillings, the strength is reduced significantly. Where no rock
wall contact exists, the conditions are extremely unfavourable to
tunnel stability.
In fractured ground, the orientation of the joints is an essential
parameter. In such cases, it is very important that the
parameters Jr and Ja should be related to the joint surface most
likely to allow failure to initiate.

 The third quotient (Jw/SRF) represents the active stress situation.


This third quotient is the most complicated empirical factor and
should be given special attention, as the SRF is defined for four
different conditions:
 stress influence in brittle blocky and massive ground
 stress influence in deformable (ductile) rock masses
 stress influence weakness zones
 stress influence in swelling rock.

The parameter Jw is a measure of water pressure, which has an


adverse effect on the shear strength of joints due to a reduction in
effective normal stress.

When the Q-system is used to determine the GSI of a rock mass,


and from that, strength according to the Hoek-Brown failure
criterion, the third quotient is not assessed, since the GSI is
based solely on fundamantal rock mass properties, not the
additional effects of water and active stress.

2.5.1 Updates to Q system


The original 1974 system was updated via further case studies by
Barton and Grimstad (1994). The main changes relate to the “stress
reduction factor” (SRF) and the “excavation support ratio” (ESR) and the
numerical values associated with these terms. In the most extreme
cases of high stress and hard massive (unjointed) rock, the maximum
SRF-value has to be increased from 20 to 400 in order to give a Q-value
which correlates with the modern rock support.

The updated values for SRF are reproduced in Appendix B.

Peck (2000) determined the Stress Reduction Factor in highly stressed


jointed rock for Australian geomechanics applications. No criteria are
given for making a choice within the range of 5 to 50 but, but if the SRF

2-11
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

rating changes from 5 to 50, the overall Q-value changes by an order of


magnitude! Therefore Peck suggested an interpolation of SRF-values
within these wide limits.

The updated values for SRF (for Australian competent rocks with rock
stress problems) are reproduced in Appendix B.

The first two quotients RQD/Jn and Jr/Ja are often used in a stope design
method in the mining industry, but their representation of „relative block
size‟ and „inter-block shear resistance‟ are not sufficient descriptions of
the degree of instability. In cases where this approach is used (i.e. the
term Jw/SRF is omitted), then the classification is referred to as Q’.

2.5.2 Application - support and reinforcement


In relating the value of the index Q to the stability and support
requirements of underground excavations, an additional parameter, the
„Equivalent Dimension‟, De, of the excavation is used. This dimension is
obtained by dividing the span, diameter or wall height of the excavation
by a quantity called the Excavation Support Ratio, ESR. Hence:

Excavation span, diameteror height


De 
ESR

The value of ESR is related to the intended use of the excavation


and to the degree of security which is demanded of the support system
installed to maintain the stability of the excavation. The values are given
in the following table.

Excavation
Description ESR
category
A Temporary mine openings. 3-5
Permanent mine openings, water tunnels for
B hydro power (excluding high penstocks), pilot 1.6
tunnels, drifts headings for large excavations.
Storage rooms, water treatment plants, minor
C road and railway tunnels, surge chambers, 1.3
access tunnels.
Power stations, major road and railway
D tunnels, civil defence chambers portal 1.0
intersections
Underground nuclear power stations railway
E 0.8
stations, sports and public facilities, factories

Barton provided additional information on rockbolt length, maximum


unsupported spans and roof support pressures to supplement the

2-12
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

support recommendations. The length L (m) of rockbolts can be


estimated from the excavation width B (m) and the Excavation Support
Ratio ESR:

0.15B
L  2
ESR

The maximum unsupported span (m) can be estimated from:

Max unsupported span  2.ESR.Q0.4

Based upon analyses of case records, the relationship between the


value of Q and the permanent roof support pressure Proof (kPa) is
estimated from:

200 J n .Q 0.333 200.Q 0.333


Proof  Proof 
3 Jr Jr

For Jn < 6 For Jn > 6

The support factor of Safety (FOS) can then be defined as;

Fb
FOS  2
Proof xSb

Where Fb refers to the bolt (tension) capacity (kN) and Sb the bolt
spacing (m).

Support and reinforcement guidelines have also been developed from


the Q system. In 1993 Grimstad proposed a summary graph of
recommendations (based Barton‟s previously tabulated version) on
different combinations or rock quality, Q, and on Equivalent Span. A
version of this graph is displayed overleaf . It must be remembered that
this graph was developed for permanent support in civil tunnels, shafts
and caverns. These recommendations are likely to be rather
conservative for many mining excavations.

2-13
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

To cover the cases for vertical walls the following adjustments were
recommended:

Qw = 5xQ for Q>10; Qw = 2.5xQ for 0.1<Q<10; Qw = 5 for Q<0.1

Where Qw is the Q value for the wall, and Q is the original Q value.

2.5.3 Strengths and Limitations


Especially in the poorer rock class (Q<1) the system may give
erroneous design outcomes. The true nature of the rock mass that is
essential for the determination of the support measures (e.g. swelling,
squeezing or popping ground) is not explicitly considered in the Q
system. For the conditions in faults and weakness zones, the supports
should be checked or designed by complimentary engineering methods.

From the mining engineer‟s perspective, even such a simple case as


block instability is much more complicated than can be assessed by a
single number like a Q value. The Q system can however be used as
an indicator for rock support or other types of rock engineering
classification.

2-14
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

2.6 Geological Strength Index (GSI)


As will be discussed in following Topics, the strength of a jointed rock
mass depends on the properties of the intact rock pieces and also upon
the freedom of these pieces to slide and rotate under different stress
conditions. This freedom is controlled by the geometrical shape of the
intact rock pieces as well as the condition of the surfaces separating the
pieces. Angular rock pieces with clean, rough discontinuity surfaces will
result in a much stronger rock mass than one which contains rounded
particles surrounded by weathered and altered material.

The Geological Strength Index (GSI) provides a system for estimating


the reduction in rock mass strength for different geological conditions.
This system is presented in the table on the next page.

One of the practical issues with using the chart is to know what is the
scale of the block size. The recommended approach is to assess the
structure on the scale of the excavation or slope under consideration.
Also, material properties will exhibit distributions about the “mean” value,
and that the distributions can have a significant impact on the design
values. Therefore, a range of values rather than a single discrete value
and answer should be considered.

Alternatively, GSI can be estimated from the Q-system (EXCLUDING


the quotient (Jw/SRF) (i.e. in the form of Q‟) as follows:

GSI  9 ln Q' 44

Where;

RQD J r
Q  x
Jn Ja

2-15
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

Slickensided, highly weathered surfaces with compact

Slickensided, highly weathered surfaces with soft clay


Smooth, moderately weathered and altered surfaces
GEOLOGICAL STRENGTH INDEX FOR
BLOCKY JOINTED ROCKS

Rough, slightly weathered, iron stained surfaces


From a description of the structure and
surface conditions of the rock mass,
pick an appropriate box in this chart.

Very rough, fresh unweathered surfaces

coatings or fillings of angular fragments


Estimate the average value of GSI
from the contours. Do not attempt to
be too precise. Quoting a range from
38 to 42 is more realistic than stating
that GSI = 38. It is also important to
recognise that the Hoek-Brown

SURFACE CONDITIONS
criterion should only be applied to
r oc k m as s es w her e t he s iz e of
individual blocks or pieces is small

coatings or fillings
c o m pa r e d w i t h t h e s i z e o f t h e

VERY GOOD

VERY POOR
e x c a v a t i o n u n de r c o n s i d e r a t i o n .
When the individual block size is
more than about one quar ter of

GOOD

POOR
the excavation size, the failure will be

FAIR
structurally controlled and the Hoek-Brown
criterion should not be used.
STRUCTURE DECREASING SURFACE QUALITY

INTACT OR MASSIVE - intact


rock specimens of massive in 90 N/A N/A N/A

DECREASING INTERLOCKING OF ROCK PIECES


situ rock with few widely spaced
discontinuities
80
BLOCKY - well interlocked un-
disturbed rock mass consisting 70
of cubical blocks formed by three
intersecting discontinuity sets
60

VERY BLOCKY - interlocked,


partially disturbed mass with 50
multi-faceted angular blocks
formed by 4 or more joint sets
40
BLOCKY/DISTURBED - folded
and/or faulted with angular blocks
formed by many intersecting
discontinuity sets 30

DISINTEGRATED - poorly inter-


locked, heavily broken rock mass 20
with mixture of angular and
rounded rock pieces

FOLIATED/LAMINATED - folded 10
and tectonically sheared. Lack
of blockiness due to schistosity N/A N/A
prevailing over other discontinuities

2-16
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

2.7 Further Applications

2.7.1 Rock Mass Modulus


Many attempts have been made to relate the classification systems
RMR and Q to the static rock mass modulus, Emass, (or modulus of
deformation) (GPa).

 Bieniawski
Emass  2RMR  100

 Barton
Emass  25log10 Q for Q>1 and generally hard rocks

 Serrafim and Pereira, where UCS < 100MPa;


 GSI 10 
c  
E mass  10  40 
100

Note that the modulus Emass is reduced progressively as the value of


σc falls below 100MPa. This reduction is based upon the reasoning
that the deformation of better quality rock masses is controlled by the
discontinuities while, for poorer quality rock masses, the deformation
of the intact rock pieces contributes to the overall deformation
process.

 Serrafim and Pereira, where UCS > 100MPa


 GSI 10 
 
E mass  10  40 

 Barton
1
 Q c  3
E mass  10  where c = UCS of intact rock material (MPa)
 100 

2-17
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

The figure below summarises these classification and modulus


relationships such that;

 15 log10 Q  40 
 
E mass  10  40 
from RMR  15log10 Q  50

However, there is not a strong correlation between the two classification


systems, and in general equations equating different classification
systems should not be used and in all cases the applicability limits of the
fitted curves must be respected.

2.8 Rock Classification Discussion

2.8.1 Advantages

 Rock Mass quality can be assessed simply, rapidly and


continuously.
 The classification values can be established by trained site
personnel (i.e. a high level of general engineering expertise is not
required).
 Continuous rock mass assessment using logging sheets will alert
mine workers and engineering staff to significant changes in rock
conditions.
 Engineering design is coherently based on previous experience.

2-18
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

2.8.2 Disadvantages

 The systems currently in use are historical and idiosyncratic and


the algebra and rating values of the systems have not been
scientifically considered.
 They cannot be used for the full range of engineering objectives.
 None of the techniques have any scientific foundation; even the
algebra and rating values have not been scientifically considered.
The systems are good as guides only and should be considered
in this manner. There are also immediate shortcomings in both:
 Stress is not included in RMR; and
 Intact rock strength in not included Q.
 They should not be used as a sole measure of the quality of the
rock mass, or as a sole guide to reinforcement except in
conceptual studies where very limited data exist.
 In both the RMR and the Q systems (and especially in the Q‟
system), there is an over-reliance on the RQD values, as
illustrated below.

 There is poor correlation between Q and RMR and even though


many relationships are presented. In general, you should not
use the many equations presented in the literature to obtain
values for one classification system based on rating obtained
using another.

2-19
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

 The Q System is unable to quantify an accurate relationship


between the geometry of the rock mass and that of the
excavation.

References
Barton, N. (2002). Some new Q-value correlations to assist in site characterization
and tunnel design. Int. Jnl Rock Mech Min Sc, 39, pp185-216
Barton, N., Lien, R., Lunde, J., (1974). Engineering classification of rock masses
for the design of rock support. Rock Mech. 6, pp189-236.
Barton, N and Grimstad, E (1994). The Q-System following twenty years of
application in NMT support selection, in Felsbau 12(6):428-436.
Bieniawski, Z.T. (1988). The rock Mass Rating (RMR system (Geomechanics
classification in Engineering Practice), in Rock Classification for Engineering
Purposes, Kirkaldie L (Ed) American Society for Testing and Materials
(ASTM), Philadelphia, pp 17-34
Hoek, E. Rock Engineering. Course Notes. www.rocscience.com
Palmstrom, A, Milne, D & Peck, W. (2000). The reliability of rock mass
classification used in underground excavation and support design. GeoEng
Workshop www.ausmin.com.au
Peck, W. (2000). Determining the Stress Reduction Factor in Highly Stressed
Jointed Rock. Australian Geomechanics, 35/2 pp57-60
www.ausmin.com.au
Stille H. & Palmstrom A. (2002). Classification as a tool for rock engineering,
Tunnelling and Underground Space Technology, 18, pp331-345

2-20
Version 1.10:Jan 2010: UNSW: GS
Appendix A –RMR System tables
MINING GEOMECHANICS

Appendix B – Q System tables

2-22
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

2-23
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

2-24
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

2-25
Version 1.10:Jan 2010: UNSW: GS
MINING GEOMECHANICS

1994 update of SRF

6 (b) Competent rock, rock stress c/ 1 / 1 SRF


problems
L Moderate slabbing after >1 hour in massive rock 5 to 3 0.5 to 0.65 5 to 50

M Slabbing and rock burst after a few minutes in 3 to 2 0.65 to 1 50 to 200


massive rock
N Heavy rock burst (strain-burst) and immediate <2 >1 200 to 400
dynamic deformations in massive rock

Australian competent rock masses (Peck, 2000)

6(b) Rock stress problems, competent c/ 1 / 1 SRF


rock, Australia
H* Low stress, near surface, open joints >200 <0.01 2.5

I* Medium stress, favourable stress condition 200 to 10 0.01 to 0.3 1

J* High stress, very tight structure. Usually 10 to 5 0.3 to 0.4 0.5 to 2


favourable to stability, may be unfavourable for
wall stability
K High stress, jointed rock. RQD/Jn<30. If 3 is 10 to 1.5 0.3 to >1 From either [3]
known then SRF = 34(c/ 1) [3]
-1.2
else if 3 is known
else [2]
SRF = 34(1/ 3) (c/ 1)
0.3 -1.2
[2]
L* Moderate slabbing after >1 hour in massive rock. 5 to 3 0.5 to 0.65 5 to 9
RQD/Jn30
# to #
M* Slabbing and rock burst after a few minutes in 3 to 2 0.65 to 1 9 15
massive rock RQD/Jn30
# #
N* Heavy rock burst (strain burst) and immediate <2 >1 15 to 20
dynamic deformations in massive rock.
RQD/Jn30
# = SRF value subject to further review and should be used with caution
* = Except for case K, correction should be made for strongly anisotropic virgin
stress field (if measured): when 1/ 3= 5 to 10, then reduce c to 0.75 c ;
when 1/ 3>10 then reduce c to 0.5 c

1994 update of ESR

Type of Excavation ESR


A Temporary mine openings. 2 to 5

B Permanent mine openings, water tunnels for hydro power (excluding high pressure 1.6 to 2.0
penstocks), pilot tunnels, drifts and headings for large excavations.
C Storage rooms, water treatment plants, minor road and railway tunnels, surge chambers, 1.2 to 1.3
access tunnels.
D Power stations, major road and railway tunnels, civil defence chambers, portal intersections. 0.9 to 1.1

E Underground nuclear power stations, railway stations, sports and public facilities, factories. 0.5 to 0.8

2-26
Version 1.10:Jan 2010: UNSW: GS
Rock
MINING GEOMECHANICS

Mechanics

3
TOPIC 3: ROCK MASS
BEHAVIOUR – INTACT
ROCK STRENGTH

3.1 Introduction
A failure criterion for a rock is a mathematical expression that defines
the stress state at which the rock will fail, and is usually expressed in
terms of the stress tensor and the material properties of the rock. The
term „failure‟ implies that the rock has completely disintegrated. It is,
however, possible for the rock to become unserviceable, in an
engineering sense, if substantial plastic strains have developed. In this
context the term „yield criterion‟ is often adopted.

Failure criteria can be;

 developed fundamentally from the mechanical analysis of an


assumed failure mechanism, or
 they can be developed empirically by modelling the observed
behaviour of rock during laboratory and/or site tests.

In some problems, it may be the behaviour of the intact rock material


that is of concern. For instance;

 the excavation of rock by drilling and blasting


 the stability of excavations in good quality, brittle rock which is
subject to rockburst conditions.

In other instances, the behaviour of single discontinuities, or of a small


number of discontinuities, will be of importance. For instance;

Version 1.10:Jan 2010: UNSW: GS 3-1


MINING GEOMECHANICS

 the stability of blocks of rock formed by the intersections of three


or more discontinuities in the roof or wall of an excavation,
 the potential for slip on a major fault.

Which of the possible yield criterion options will apply in a given case will
depend on the size of the excavation relative to the discontinuity
spacing, the imposed stress level, and the orientations and strengths of
the discontinuities, as shown below.

Remember that the presence of major discontinuities or of a number of


joint sets does not necessarily imply that the rock mass will behave as a
discontinuum. Where the rock surrounding the excavation is always
subject to high compressive stresses, it may be reasonable to treat a
jointed rock mass as an equivalent elastic continuum. Rock material and
discontinuity properties may be combined to obtain the elastic properties
of the equivalent continuum (rock mass) by using either RMR or Q-
system approached and the associated equations for estimating Emass
(as outlined in Topic 2).

In this topic, we will focus on two fundamental failure criteria for intact
rock, the Coulomb criterion and the Hoek-Brown criterion. Failure
criteria for discontinuities and the rock mass as a whole will be
discussed in Topics 4 and 5.

Version 1.10:Jan 2010: UNSW: GS 3-2


MINING GEOMECHANICS

3.1.1 Learning Objectives


By the time you have completed this Topic you will be able to:

1. Explain the basis for the Coulomb and Hoek & Brown failure
criteria in predicting intact rock failure.
2. Calculate the strength of an intact rock mass following
Coulomb or Hoek-Brown failure criteria.
3. Analyse basic laboratory rock testing results to determine the
strength of an intact rock sample following Coulomb or Hoek-
Brown failure criteria.
4. Determine the equivalent Coulomb instantaneous shear
strength parameters Hoek-Brown criterion at specific effective
normal stress values.

3.2 Review – Analysis of Stress 1

The theory has already been developed in earlier Units and in Soil
Mechanics Topic 7. The following section has been imported, with
minor modifications, from Sections 1, 4 & 6 of the Deformable Materials
Module (Priest, 2010) in order to provide relevant theoretical
background, and are reproduced here with permission.

3.2.1 Introduction to stress


In its broadest sense, stress is force per unit area. In rock mechanics
applications the units kN/m2 (kilonewtons per square metre) and MN/m2
(Meganewtons per square metre) or kPa (kilopascals) and MPa
(Megapascals) are usually adopted for stress (1 pound per square inch
= 6.895 kPa, 1 MPa = 145.0 psi). If a uniformly distributed force N acts
normal to a surface of area A, the normal stress σ is given by;

N

A

The normal stress value σ is sometimes referred to as the total stress.


This terminology is adopted because there might be a water pressure u
acting across the surface. The effective normal stress σ' acting on the
surface is given by;

 '   u

1
Section 3.2 has already been covered in the „Deformable Materials‟ Module and are presented here
if you require them for revision and continuity between Modules.

Version 1.10:Jan 2010: UNSW: GS 3-3


MINING GEOMECHANICS

The term „effective normal stress‟ is adopted because the stress carried
by water pressure plays no part in mobilising frictional shear strength on
the surface; it is only the effective stress developed by grain to grain
contacts that serves to mobilise frictional shear strength. Although the
effective stress principle is absolutely crucial in saturated and partially
saturated soils, which have an interconnected pore matrix, the principles
of effective stress are not applicable to all rock masses. Crystalline rock
materials are effectively impermeable and do not have an
interconnected pore matrix that allows water pressure to exert its effect.
Moreover, the presence of a fracture network, which is usually
responsible for transporting water through the rock mass, will apply a
further level of complexity to the problem. Throughout this document,
although the dash convention has been omitted for brevity, the stress
and strength parameters can be taken as effective stress parameters.

If a uniformly distributed force S acts parallel to a surface of area A, the


shear stress τ is given by;

S

A

There is no need to specify total and effective shear stress values


because water cannot transmit shear stresses. Shear stress levels are,
therefore, independent of the hydrostatic pressure.

The normal and shear forces within rock masses vary from place to
place, so the assumption of a uniformly distributed force will only be
valid if the area A of the surface under consideration is very small.
Under these conditions the stress is referred to as infinitesimal stress
and can be considered to be uniform across the area under
consideration.

3.2.2 Co-ordinate system convention


A right-handed Cartesian* (orthogonal) coordinate system will be used
throughout, in which;

 The positive direction of the X axis is horizontal to the north


 The positive direction of the Y axis is horizontal to the east
 The positive direction of the Z axis is vertically down

[*Cartesian is the latinised form of the surname of René Descartes, French


philosopher and mathematician (1596–1650). We recognise this etymology by
using the capital „C‟.]

Version 1.10:Jan 2010: UNSW: GS 3-4


MINING GEOMECHANICS

X horizontal north

Vx

Vy
a
b Y horizontal east
g
V

Vz

Z vertical down

This XYZ Cartesian coordinate system reflects global geographical


directions, so it is usually referred to as a Global Cartesian coordinate
system. The Cartesian coordinate system is also right-handed. The
„handedness‟ of a Cartesian coordinate system relies on an ordering of
the axes (X,Y,Z, or L,M,N or 1,2,3 etc). If we associate;

 the first axis (X, L or 1 etc) with our index finger;


 the second axis (Y, M, 2 etc) with our middle finger;
 and the third axis (Z, N, 3 etc) with our thumb.

we can create a representation of a Cartesian coordinate system with


our hand by positioning these digits at right angles to each other. If the
physical alignment of the axes corresponds to our right hand, then we
have a right-handed system. A right-handed system can be converted
to a left-handed system by reversing the direction of any one axis; for
example the above system would be left-handed if the Y axis were
changed to an orientation of horizontal to the west.

The best way to remember it, is that the fingers of the right hand point
from the Vx axis to the Vy axis, and if the thumb of the right hand then
points in the positive direction of the Vz axis, then the system is right-
handed. If the Vz axis points in the opposite direction, then the system is
left-handed.

We normally work exclusively with right-handed Cartesian coordinate


systems. Note that if we face north, then look down along the Z axis,
the X axis is „vertical‟ and the Y axis is „horizontal‟, which is the opposite
way round from their „usual‟ orientation on two-dimensional graphs.

Version 1.10:Jan 2010: UNSW: GS 3-5


MINING GEOMECHANICS

This minor anomaly is caused by adopting the convention „positive Z


down‟. Although the global coordinate system defined above will be
used for all formal calculations in these notes, some authors adopt
coordinate systems with the Z axis vertically up to enhance the visibility
of three-dimensional sketches. It is sometimes convenient to orientate
our coordinate axes in alignment with our local mine coordinates, some
existing geological structure, our excavation geometry or indeed our
measured stress directions. In this case we describe the coordinate
system as a Local Cartesian coordinate system, and if feasible label the
axes LMN, or some nomenclature other than XYZ, to emphasise the
local nature of these axes. This concept will be expanded in later
Topics.

3.2.3 Three-dimensional stress


To analyse three-dimensional stress it is convenient to consider the
stresses produced by forces acting on the faces of a very small cubic
element within the material. It is usual to orientate the corners of the
cube so that they are parallel to the specified axes of a Cartesian
coordinate system. Since the faces of the cube are very small it can be
assumed that the stress values do not change significantly from one
side of the cube to the other.

X
zz
zx
zy
Y
yz
yy
xy 
yx

xz
xx

The three faces of the cube that each have a corner touching the origin
of the coordinate system are here referred to as „origin faces‟; the other
three faces are referred to as „obverse faces‟. Each face can be
identified by the coordinate axis that runs normal to it. So in the Figure
above, we can see the origin-x, origin-z and obverse-y faces.

On each face of the stress cube there are force components acting in
the X, Y and Z directions. Dividing these force components by the area
of the face produces three stress components on each face of the cube,

Version 1.10:Jan 2010: UNSW: GS 3-6


MINING GEOMECHANICS

giving a total of 18 stress components. The positive directions of nine of


these stress components, acting on the three visible faces, have been
drawn and labelled in the Figure.

It is important to spend some time understanding the nomenclature and


sign convention. Taking a typical stress component xy ;

 the Greek symbol „sigma‟ indicates this is a stress component.


 the first subscript x indicates this stress component acts on a
plane whose normal is in the x direction.
 the second subscript y indicates this stress component acts
along the y direction.

The stress components xx, yy and zz act along the normals to their
respective faces and are referred to as normal stresses. For brevity
the second subscript is sometimes omitted, so the normal stresses
become x, y and z.

The other stress components, such as xy yz etc. act parallel to their
respective faces and are referred to as shear stresses. To emphasise
the shearing nature of these stresses the Greek symbol „sigma‟ is often
replaced by the Greek symbol „tau‟ to give components xy yz etc.

The sign convention for normal stresses is:

Compressive normal stresses are positive.

The logic in taking compressive stresses as positive in rock mechanics


is that almost all stresses in rock are compressive, adopting the
mechanical engineering convention would necessitate a pointless
repetition of negative signs. So, for example, normal stress of 1.45 MPa
is taken to be compressive; a normal stress of –6.75 MPa is tensile.

The sign convention for shear stresses is:

(a) If the positive normal stress on a given face acts in the


positive direction of the coordinate axis, then positive shear
stresses on this face act in the positive directions of the
relevant coordinate axes.

(b) If the positive normal stress on a given face acts in the


negative direction of the coordinate axis, then positive shear
stresses on this face act in the negative directions of the
relevant coordinate axes.

Case (a) applies to the three origin faces of the stress cube; case (b)
applies to the three obverse faces of the stress cube.

Version 1.10:Jan 2010: UNSW: GS 3-7


MINING GEOMECHANICS

In order to satisfy static equilibrium of the stress cube, ie such that it


does not displace or rotate, we require that;

xx,origin = xx,obverse

yy,origin = yy,obverse

zz,origin = zz,obverse

Each of the three groups of two normal stresses above can be regarded
as a single normal stress. The origin and obverse components are of
equal magnitude but act in opposite directions.

Also;

xy,origin = yx,origin = xy,obverse = yx,obverse

yz,origin = zy,origin = yz,obverse = zy,obverse

xz,origin = zx,origin = xz,obverse = zx,obverse

Each of the three groups of four shear stresses above can be regarded
as a single shear stress. The origin and obverse components are of
equal magnitude but act in opposite directions to ensure static
equilibrium against rotation of the stress cube.

These equalities mean that there are only six independent stress
components, which define the stress matrix, as follows

 xx  xy  xz 
  yx  yy  yz 

  zy  zz 
 zx

In view of the equalities outlined above, the stress matrix is symmetrical,


specifically xy = yx, yz = zy and xz = zx. Although we have taken
care in defining our nomenclature, these equalities mean that the
ordering of the subscripts for our shear stresses is not really important.

If a shear stress xy is computed to have a value of, say, 2.73 MPa then
the actual directions of action of the four components (xy,origin = yx,origin =
xy,obverse = yx,obverse) are the same as the positive shear stress
directions drawn on the stress cube. If a shear stress xy is computed to
have a value of, say, 3.85 MPa then the actual directions of action of
the four components are in the opposite direction to the positive shear
stress directions drawn on the stress cube. This property of shear
stresses is one of the more confusing, and least understood, aspects of
stress analysis.

Version 1.10:Jan 2010: UNSW: GS 3-8


MINING GEOMECHANICS

3.3 Coulomb Failure Criterion


Some of the basic concepts described above are now used to
describe one of the most widely applied fundamental failure criteria
for rocks, the Coulomb criterion. The following section has been
imported, with minor modifications, from Section 13 of the
Deformable Materials Module (Priest, 2010) in order to provide
relevant theoretical background, and are reproduced here with
permission.

This criterion assumes that rock failure occurs as a


result of shear displacement along a shear surface
cutting through the intact rock. This is quite a
sensible assumption since we can observe shear
failure in laboratory specimens and also at a large
scale in rock masses.

Shear failure is predicted to occur when the shear stress on a particular


surface exceeds the shear strength f, where;

 f   n tan   c

where f = shear strength (Nm-2)


n = normal stress component across incipient failure
plane (Nm-2)
c = cohesion (Nm-2) which is the shear strength at zero
normal stress, is related to the cementing between
the rock grains and crystals on the shear surface.
 = angle of friction (°) which is the frictional strength
mobilised as a result of the normal stress acting
across the shear surface.

In terms of intact rock, the Coulomb theory postulates that a material will
fail when the shear stress on the failure surface (inclined at angle β) has
reached the shear strength of the material, which is dependent on the
normal stress acting on the surface. In tension, however, the material

Version 1.10:Jan 2010: UNSW: GS 3-9


MINING GEOMECHANICS

will fail when the tensile or principal stress has reached a limiting value.
Thus failure of the rock may either occur by shear or tensile failure.

b

The Figure shows a two-dimensional view of the normal stress n and


the shear stress  acting on an inclined plane cutting through a
cylindrical or prismatic specimen of rock subjected to an axial principal
stress 1 and a lateral or confining stress 3. The normal to the inclined
plane makes an angle β to the 1 axis.

Development of the Coulomb criterion is based on a two-dimensional


stress transformation which resolves to give;

 1  3   1  3 
n   cos2b
2 2

and

 1  3 
 sin2b
2

At the point of failure the shear strength is balanced by the shear stress,
ie f = . Substituting for n and f gives;

2c   3 sin 2 b  tan  1  cos 2 b 


1 
sin 2 b  tan  1  cos 2 b 
or

2c   3 tan  
1   3 
 tan  
sin 2b 1  
 tan b 

If there is no discontinuity cutting through the specimen, or if the strength


and/or orientation of the discontinuity are such that shear failure does

Version 1.10:Jan 2010: UNSW: GS 3-10


MINING GEOMECHANICS

not occur along the discontinuity plane, then when 1 reaches a


sufficiently high level the specimen will theoretically fail by creating a
shear surface through intact rock. The rock will theoretically „choose‟ the
failure surface that gives the minimum strength according to the
previous equations. The minimum value for 1 is given by;

 tan  
 1min  sin 2b 1  
 tan b 

There is a critical plane on which the available shear strength is first


reached as σ1 increases. Inclination of this plane is given by;


bcritical  45 
2

For the critical plane, at yield;

2c cos   3 1  sin  
1 
1  sin 

Plotting 1 against 3 gives a straight-line relationship as shown below,


where the region below the line represents a stable stress state. The
vertical-axis intercept of this straight line gives the theoretical value of 1
at which the rock specimen will fail when 3 = 0, which is of course the
uniaxial compressive strength (UCS) of the rock c.

Thus;

2c cos 
c 
1  sin 

The slope of the line is given by;

1  sin  
tan   tan 2  45  
1  sin  2

The uniaxial tensile strength t of the rock is, theoretically, given by

2c cos 
t 
1  sin 

Version 1.10:Jan 2010: UNSW: GS 3-11


MINING GEOMECHANICS

Finally, the value for σ1 at failure can be expressed in the following


alternative compact form;

1  2c tan   3 tan

We can construct the Mohr-Coulomb envelope via the results of the


uniaxial compressive test and the uniaxial tensile test (or alternately
Brazilian tensile test)

 c
tan 
n
 

Compressive forces are positive so the Mohr circle for the UCS plots to
the right of the shear axis (beginning at the origin) and the UTS plots to
the left (also beginning) a the origin. The failure locus is defined from
the tangent to the two circles. In practice, extrapolating the line to the
left of the axis is fraught with difficulty. When it is satisfactorily
measured, it takes values that are generally lower than those predicted.

For this reason a tensile cut-off is usually applied and σt = 0.

In triaxial compression testing, individual core samples are subjected


initially to confinement stress then are axially loaded to failure. The
failure axial and lateral stress conditions for each test are recorded (1,
3) and are plotted as biaxial stress circle points on the normal stress
axis of a shear stress/normal stress plot, as illustrated in the following
diagram. Principal (biaxial) stress circles (Mohr circles) are constructed

Version 1.10:Jan 2010: UNSW: GS 3-12


MINING GEOMECHANICS

to join each of the stress pair intercepts on the normal stress axis by
assuming that the centre of each circle exists at a position = ½(σl + σ3)
and that each circle has a radius equivalent to ½(σl - σ3). The failure
locus is generated by drawing the best fit line tangent to all of the stress
circles which are plotted, as shown:

 c
tan 
n
 

Comments on Coulomb Failure Analysis


The Coulomb failure criteria is generally not satisfactory because:
 Implies major shear fracture at peak strength. This is not always
the case.
 Implies a direction of shear failure which does not always agree
with experimental observations.
 Experimental peak strength envelopes are generally non-linear.
They can be considered linear over limited ranges of n or 3

Version 1.10:Jan 2010: UNSW: GS 3-13


MINING GEOMECHANICS

3.4 The Hoek - Brown Failure Criterion


The following section has been imported, with minor modifications, from
Section 14 of the Deformable Materials Module (Priest, 2010) in order to
provide relevant theoretical background, and are reproduced here with
permission.

The empirical failure criterion is derived from experimenting with


distorted parabolic (best fit) curves, based on observations of
biaxial/triaxial principal stress states and unconfined compression failure
conditions of rock.

The Hoek-Brown criterion has been widely adopted in rock mechanics


applications for the following reasons:

 The Hoek-Brown criterion was developed specifically for rock


materials and rock masses.
 Input parameters for the Hoek-Brown criterion can be derived
from uniaxial testing of the rock material, mineralogical
examination, and measurement of the rock mass fracture
characteristics.
 The Hoek-Brown criterion has been applied for some 30 years by
practitioners in rock engineering, and has been applied
successfully to a wide range of intact and fractured rock types.

The locus of rock failure follows a path which can be defined in terms of
biaxial principal failure stresses and the rock unconfined compressive
strength (i.e.- compression failure model). The generalized form of the
Hoek and Brown criterion is;
a
  
 1   3   ci  mb 3  s 
  ci 

And for intact rock simplifies to;


0.5
    1   3  mi c i 3   Ci2
 1   3   ci  mi 3  1 or
  ci 

Where

1 = major principal effective stress at failure;

3 = minor principal effective stress at failure;

Version 1.10:Jan 2010: UNSW: GS 3-14


MINING GEOMECHANICS

 ci = UCS of the intact rock material;

mb = material constant for the rock mass.

mi = material constant for the intact rock.

s,a = constant, related to rock mass characteristics

For intact rock, values of mi (the Hoek-Brown constant m for intact rock
material), depends on the rock type and mineralogy. Typical values are
shown below after Marinos and Hoek (2000).

Version 1.10:Jan 2010: UNSW: GS 3-15


MINING GEOMECHANICS

Since the relationship between the principal stresses at failure for a


given rock is defined by two constants (s=1, a=0.5 for intact rock), the
uniaxial compressive strength σci and a constant mi , lab tests may also
be used to obtain these values.

The range of minor principal stress values over which these tests are
carried out is critical in determining reliable values for the two constants.
A range of 0 < σ3 < 0.5σci should be used in any laboratory triaxial tests
on intact rock specimens. At least five well spaced data points should
be included in the analysis.

Characteristic rock stress versus strength curves for the Hoek and
Brown failure condition are developed using a plot of major principal
stress components along the vertical axis and minor principal stress
components along the horizontal axis, as illustrated below (after Hoek
and Brown 1980).

Once five or more triaxial test results have been obtained, they can be
analysed to determine the uniaxial compressive strength σci and the
Hoek-Brown constant mi. Ideally the values of the constants should be
determined by statistical analysis with a coefficient of determination (r2)
in excess of 90%:

y  m ci x  s ci

where

Version 1.10:Jan 2010: UNSW: GS 3-16


MINING GEOMECHANICS

y   1   3  and x   3
2

3.4.1 Relating Hoek-Brown Intact Rock Strength to


Coulomb Strength Parameters
Most of the currently used techniques for analysing the stability of near
surface structures, such as rock slopes, are based on the application of
the Coulomb shear strength parameters cohesion c, and the angle of
friction Φ on an anticipated shear surface subjected to a normal stress
σn. If the Hoek-Brown criterion is to be used to model the strength of
near surface fractured rocks, it is necessary to determine equivalent
Coulomb shear strength parameters for the specified level of effective
normal stress. Since the Hoek-Brown criterion is non-linear, the values
of cohesion and friction will vary with the value of the effective normal
stress at which they are computed. It is desirable, therefore, to refer to
these equivalent Coulomb parameters as „instantaneous‟ values ci and
Φi adding the subscript “i” to emphasise that they are applicable at one
specified normal stress level. Do NOT confuse this subscript with the „i‟
implying intact rock conditions. For intact rock, equivalent cohesion and
friction parameters for the Hoek-Brown criterion for a shear surface
subjected to a normal stress can be determined from the following
expressions (for intact rock only);

Letting σ3 = r σci gives, for the Hoek-Brown criterion;

 1HB  r ci   c rmi  1

and for the Coulomb criterion;

 1MC   ci  r ci  tan

Equating the expressions and simplifying (tanψ = tan2((π/4) + (Φi/2)) we


obtain;

  r  1  rmi  1   
  2arctan  
  r  4
   

and hence cohesion can be found from;

 c 1  sini 
c
2 cosi

Version 1.10:Jan 2010: UNSW: GS 3-17


MINING GEOMECHANICS

The above method for calculating the equivalent angle of friction and
cohesion for a Hoek-Brown material preserves the „correct‟ value for the
Coulomb uniaxial compressive strength, so it can be referred to as the
„uniaxial compressive strength‟ strategy.

In Topic 5, we will develop a similar approach, using the „angle of


friction‟ strategy to determine instantaneous cohesion and friction for the
general case of a rock mass.

References
Priest S. D. (2010). Deformable Materials Module lecture notes, Mining
Geomechanics. The University of Adelaide.

Version 1.10:Jan 2010: UNSW: GS 3-18


Rock
MINING GEOMECHANICS

Mechanics

4
TOPIC 4: ROCK MASS
BEHAVIOUR –
DISCONTINUITY
STRENGTH

4.1 Introduction
As we already know, the two main mechanical components of a rock
mass are:

1. the rock material (or intact rock)


2. the defects (or discontinuities)
In Topic 3 we covered intact rock material yield models. In this Topic
we will examine the shear behaviour of the rock defects (discontinuities).
The combined properties will be addressed in Topic 5 (Rock Mass
Strength and Behaviour). Having stated that shear behaviour is one
part of rock mass behaviour, it must also be emphasized that shear
behaviour of rock defects can be a controlling mechanism alone. The
stability of open pit slopes may be strongly influenced by the presence of
major defects (such as bedding planes or faults) dipping out of the face
and thus the slope stability will be calculated on the basis of the shear
behaviour of the defects. In the underground situation, dilution in an
open stope may be affected by the presence of local defects in the
hangingwall, or defect infill may influence the stability of the backs.

All rock masses contain discontinuities such as bedding planes, joints,


shear zones and faults. At shallow depth, where stresses are low,
failure of the intact rock material is minimal and the behaviour of the rock
mass is controlled by sliding on the discontinuities. In order to analyse
the stability of this system of individual rock blocks, it is necessary to

Version 1.10:Jan 2010: UNSW: GS 4-1


MINING GEOMECHANICS

understand the factors that control the shear strength of the


discontinuities which separate the blocks.

4.1.1 Learning Objectives


By the time you have completed this topic you will be able to:

 Describe the typical shear resistance-displacement behaviour of


smooth and rough joints.
 Explain the effect of increasing normal stress on above joint
behaviour.
 Describe the effect of surface roughness on joint shear strength
and explain the concept of roughness angle and its effect on
initial shear resistance according to Patton.
 Explain the effect of roughness on dilation during shear on a joint
and its effect on rock stability in confined and unconfined
situations.
 Describe qualitatively the effect of infilling on joint strength.
 Explain direct shear testing methods and data analysis required
to estimate shear strength.

4.2 Shear Strength of Planar Surfaces


Suppose that a mine site sends a number of samples of rock to a
commercial testing laboratory for shear testing. Each sample contains a
perfectly planar (absolutely planar, having no surface irregularities or
undulations), through-going (continuous) bedding plane that is
cemented. Each specimen is subjected to a stress σn normal to the
bedding plane, and the shear stress τ, required to cause a displacement
δ.

The resulting behaviour is illustrated on the next page. Shear stress will
increase rapidly until the peak strength is reached, and as the

Version 1.10:Jan 2010: UNSW: GS 4-2


MINING GEOMECHANICS

displacement continues, the shear stress will fall to some residual value
that will then remain constant.

Repeating the test at various normal stresses and plotting the peak and
residual shear strengths for these different normal stresses results in the
two lines illustrated below. For planar discontinuity surfaces the
experimental points will generally fall along straight lines. The peak
strength line has a slope of  and an intercept of c on the shear strength
axis. The residual strength line has a slope of r.

The relationship between the peak shear strength Tf and the normal
stress σn can be represented by the Coulomb strength criterion:

 f   n tan   c

where f = shear strength (Nm-2)


n = normal stress component across incipient failure
plane (Nm-2)
c = apparent cohesion (Nm-2)
 = apparent angle of friction (°)

In the peak strength plot above, the line extended back to the vertical
axis determines the cohesion. The extension of the locus to the vertical

Version 1.10:Jan 2010: UNSW: GS 4-3


MINING GEOMECHANICS

axis is fictitious (as a clean defect would have no cohesion), so for clean
defects, consider this an „apparent friction‟, which is often the case for
„intact‟ or rock material as was discussed in the previous Topic.

4.3 Shear Strength of Rough Surfaces

4.3.1 Dilatancy Models: Patton


A sawn or ground surface is often used to determining the basic friction
angle (as described above). However, a natural discontinuity surface in
hard rock is never as smooth as this „ideal‟. The undulations and
asperities on a natural joint surface have a significant influence on its
shear behaviour. Generally, this surface roughness increases the shear
strength of the surface, and this strength increase is extremely important
in terms of the stability of excavations in rock.

Patton was the first researcher to relate the shear behaviour of joints to
normal load and roughness, based on his work of a model of a joint in
which roughness is represented by a series of constant-angle triangles
or saw-teeth, as shown below.

Version 1.10:Jan 2010: UNSW: GS 4-4


MINING GEOMECHANICS

In his model, the dilatancy angle, i, is defined as the arc tangent of the
ratio between vertical and shear displacement of the sample during the
shearing. The model assumes the rock is rigid and the dilatancy angle
constant. Patton observed;

 at low normal loads, when there was practically no shearing of


asperities, the shear strength of the joints was:

 f   n tan   i 

Where b = basic angle of friction (°)

 at high normal loads, when the tips of most asperities were


sheared off, he found reasonable agreement with experimental
results using a different failure criterion:

 f   n tan r   c

Where r = residual angle of friction (°)

Combining the two failure criteria together, Patton obtained a bilinear


envelope (in the figure below) that describes the shear strength of plane
surfaces containing a number of regularly spaced teeth of equal
dimensions.

Version 1.10:Jan 2010: UNSW: GS 4-5


MINING GEOMECHANICS

This model takes into account the effect of the asperities, however, the
criteria are not satisfactory for describing the shear behaviour of irregular
rock surfaces. Patton explains the discrepancy with real joints by
suggesting that the failure envelope for rock surfaces reflects changes in
the intensities of different modes of failure occurring simultaneously
(Graselli, 2001).

4.3.2 Dilatancy Models: Ladanyi & Archambault


Again looking at two-dimensional saw-tooth profiles, the transition from
dilatancy to shearing was studied theoretically and experimentally by
Ladanyi & Archambault (1970) who approached the problem of joint-
shear strength by identifying the areas on the joint surface where sliding
and breaking of asperities are most likely to occur. They define as to be
the area where shearing through the asperities takes place. Over the
rest of the surface (1- as ), the asperities are assumed slide over each
other without damage. The proposed expression for the total resisting
force is given as:

    
4
    
(1  )1.5  1   tan i  tan   0.232 1  (1  )1.5  (1  10 ) 0.5
 J  J   J    J  J
  
   
J
1- (1  )5.5 tan i  tan 
 J 

where

/ J = ratio of effective normal stress on the joint to the


uniaxial compressive strength of the joint wall
τ/ J = ratio of joint shear strength to the uniaxial
compressive strength of the joint wall

Version 1.10:Jan 2010: UNSW: GS 4-6


MINING GEOMECHANICS

4.3.3 Dilatancy Models: Barton’s Model


Pattons equation is valid at low normal stresses where shear
displacement is due to sliding along the inclined surfaces. At higher
normal stresses, the strength of the intact material will be exceeded and
the teeth will tend to break off, resulting in a shear strength behaviour
which is more closely related to the intact material strength than to the
frictional characteristics of the surfaces. In reality, changes in shear
strength with increasing normal stress are gradual rather than abrupt.

An alternative approach to the problem of predicting the shear strength


of rough joints was proposed by Barton (1977). Based on tests carried
out on natural rough joints, Barton derived the following empirical
equation:

  JCS  
 p   n tan  b  JRC  Log10   
   n 

where

n = normal stress

b = the basic friction angle of a smooth (saw-cut) joint


surface,

JRC = Joint Roughness Coefficient (ranges from 0 to 20)

JCS = Joint-wall Compressive Strength (equivalent to j in


previous equations).

That is, at a particular scale:

  JCS  
i   JRC log10   

   n 

JRC (Joint Roughness Coefficient) is a parameter that represents the


roughness of the joint and JCS (Joint Compressive Strength) is the
compressive strength of the rock on the joint surface, taking into account
possible reductions in resistance resulting from fatigue, chemical
alteration, or other processes that weaken the rock at the interface.
NOTE that units of both JRC, JCS and σn must be the same in this
equation.

Barton‟s equation is probably most applicable for stresses in the range;


0.01   0.3
j

Version 1.10:Jan 2010: UNSW: GS 4-7


MINING GEOMECHANICS

This is also the range into which most slope stability problems fall, so
Barton‟s equations are applicable to slope stability.

Estimates of JCS
When the joint is “fresh,” JCS is equal to the compressive strength of the
rock (i.e. JCS= c). Where joint walls are weathered to a moderate
depth, methods of point load testing and Schmidt Hammer techniques
can be used as outlined in Hoek (2000) and shown below (using
uniaxial compressive strength values for JRC). Where no direct
measurements are available, a ratio of JCS/c = 0.25 may be used.

Version 1.10:Jan 2010: UNSW: GS 4-8


MINING GEOMECHANICS

Estimates of JRC
The joint roughness coefficient JRC is a number that can be estimated
by;

 Back analysis of shear tests(if JCS is known), or


 Comparing the appearance of a discontinuity surface with
standard profiles

Back analysis of shear tests

arctan  /  n   b
JRC 
 JCS 
log10  
 n 

Back analysis via field tilt tests:

  b
JRC 
 JCS 
log10  
 n 

Note that very low normal stresses are usually involved in field tilt tests.
In a tilt test, two blocks (1 fixed to the bottom and one resting on top) are
placed on the apparatus. The blocks are tilted until top block begins to
slide. The angle at which this movement begins is read as angle 

Standard profiles have been published by Barton and others. One of


the most useful of these profile sets was published by Barton and
Choubey (1977) and shown on the next page.

The appearance of the discontinuity surface is compared visually


with the profiles shown and the JRC value corresponding to the
closest match is chosen.

Version 1.10:Jan 2010: UNSW: GS 4-9


MINING GEOMECHANICS

In the case of small scale laboratory specimens, the scale of the surface
roughness will be approximately the same as that of the profiles
illustrated. In the field the length of the joint surface may be much larger
and the JRC value must be estimated for the full scale sliding surface,
using the chart on the next page.

Version 1.10:Jan 2010: UNSW: GS 4-10


MINING GEOMECHANICS

Version 1.10:Jan 2010: UNSW: GS 4-11


MINING GEOMECHANICS

4.3.4 Practical application of dilatancy models


Note that Barton‟s criteria is also only valid fro rock-to-rock contact
and;

0   n  JCS

and the peak shear strength curves should be truncated for design
purposes at a maximum allowable strength given by;

     
arctan    70  i.e   r  JRC log10  JCS    70
   
n    n 

Barton‟s equation is in close agreement with Ladanyi and Archambault


at very low normal stress levels, however the equations diverge as the
normal stress increases. This is because Barton‟s equations reduce to
f = tan as J tends to 1; where as Ladanyi and Archambault‟s
reduces to f= R, the shear strength of the rock material adjacent to the
joint surface. Barton‟s equation therefore tends to be more conservative
at higher stress levels.

Scale effects
Surface topography of joints varies widely and features can be
distinguished into;

 Asperities (small scale roughness)


 Larger protrusions (immediate roughness)
 Undulations (large scale roughness) – as shown below relative to
sheared length D.

Effects of these broad classes is related to the sheared length. For


example, sampling/measuring inclinations i over increasing D gives an

Version 1.10:Jan 2010: UNSW: GS 4-12


MINING GEOMECHANICS

inverse relationship Now consider the results of a series of shear tests


on a single rough surface, 600mm by 600mm (shown on the next page)
which has been split into various sizes:

Referring to each „sample‟;

1. The first test examines the shear behaviour of 100mm diameter


samples – similar in size to a core sample. The shear behaviour
is determined by small bumps and ripples (i.e. Patton‟s 2nd
order projections).
2. If we group these surfaces in pairs, the surfaces are now
(roughly) 200mm in diameter. The effect of the very small
bumps and ripples becomes less pronounced
3. Now if we examine half the total surface area of the sample,
the second order bumps and ripples have almost no effect and
the behaviour is dominated by the first order waviness
4. Now finally examine the behaviour of the whole surface.

The contribution of the various components (asperity, geometric and


basic frictional) are shown below;

Version 1.10:Jan 2010: UNSW: GS 4-13


MINING GEOMECHANICS

Another effect of the scale effect is the increased mobilisation required


to reach peak strength (for any constant normal load). This means that
as the defect area increases, the defect needs to move increasingly
larger distances to achieve that value. In the figure below, the defect
depicted by „sample‟ 4 achieves peak strength after a greater
displacement than that of 1. For a given normal load, the larger the
scale the longer the defect must move before it reached peak strength.

Thus, the following scale corrections for JRC are used:


0.02 JRCo
L 
JRCn  JRCo  n 
 Lo 

And for JCS


0.03 JRCo
L 
JCSn  JCSo  n 
 Lo 

Where

JRCo & Lo: 100mm lab scale

JRCn & Ln: in situ block size

Version 1.10:Jan 2010: UNSW: GS 4-14


MINING GEOMECHANICS

4.4 Joint Infilling


Up to now, we have been discussing situations where there is rock-rock
contact across the joint. This shear strength can be reduced drastically
when part or all of the surface is not in intimate contact, but covered by
soft filling material such as clay gouge. For planar surfaces, such as
bedding planes in sedimentary rock, a thin clay coating will result in a
significant shear strength reduction. For a rough or undulating joint, the
filling thickness has to be greater than the amplitude of the undulations
before the shear strength is reduced to that of the filling material.

Consider a joint with an asperity dimension „a‟ and a filling thickness


of „t‟. As the gouge or infill increases in thickness, for a planar
surface (a = o), there will be no rock contact and shear strength
reduces to that of the infill. For non-planar surfaces;

 t << a, almost immediate rock contact, shear strength may be 10-


20% lower due to smaller  b.
 t < a, rock interaction possible after some shear displacement
 t > a, no rock contact; possible development of high pre-
pressures
 t >> a, the surface roughness ceases to exert any influence and
the shear strength is controlled by the infill alone

Important aspect is rock wall and infill interaction (if any) as rock-rock
strength is far greater than infill material strength . Published data can be
used to assist in quantifying this reduction in shear strength, by referring
to the shear strength parameters of typical rock with commonly found
filling materials as given on the next page.

Version 1.10:Jan 2010: UNSW: GS 4-15


MINING GEOMECHANICS

Rock Description Peak Residual Residual


Peak Φ°
c’ (MPa) c’ (MPa) Φ°
Basalt Clayey breccia, wide
0.24 42
variation from clay to
Bentonite Bentonite seam in chalk 0.015 7.5
Thin layers 0.09-0.012 12-17
Triaxial tests 0.06-0.1 9-13
Bentonitic shale Triaxial tests
0-0.27 8.5-29
Direct shear tests 0.03 8.5
Clays Over-consilidated, slips,
0-0.18 12-18.5 0-0.003 10.5-16
joints and minor shears
Clay shale Triaxial tests
0.06 32
Stratification surfaces 0 19-25
Coal measure rocks Clay mylonite seams, 10 to
0.012 16 0 11-11.5
25mm
Dolomite Altered shale bed, ±
0.04 14.5 0,02 17
150mm thick
Dolorite,granodiorite and Clay gouge (2% clay, PI =
0 26.5
porphyry 17%)
Granite Clay filled faults 0-0.1 24-45
Sandy loam fault filing 0.05 40
Tectonic shear zone,
schistose and broken
0.24 42
granites, disintegrated rock
and gouge
Greywacke 1 – 2mm clay in bedding
0 21
planes
Limestone 6mm clay layer
10-20mm clay fillings 0.1 13-14 0 13
<1 mm clay filling 0.05-0.2 17-21
Limestone, marl and Interbedded lignite layers 0.08 38
Lignites Lignite/marl contact 0.1 10
Limestone Marlaceous joints, 20mm
0 25 0 15-24
thick
Lignite Layer between lignite and
0.014-0.3 15-17.5
clay
Montmorillonite Bentonite 80 mm seams of bentonite 0.36 14
0.08 11
clay (montmorillonite) clay in 0.016-.02 7.5-11.5
Schists, quartzites and 100-15-mm thick clay
0.03-0.08 32
siliceous schists filling
Stratification with thin clay 0.61-0.74 41
Stratification with thick clay
0.38 31
Slates Finely laminated and
0.05 33
altered
Quartz/kaolin/pyrolusite Remoulded triaxial tests 0.042-.09 36-38

Version 1.10:Jan 2010: UNSW: GS 4-16


MINING GEOMECHANICS

4.5 Testing: Direct Shear Tests


Depending on the testing machine, direct shear tests may be conducted
upon full sized rock specimens in either lump (irregular) or prepared
cylinder forms. The specimens are failed in shear through the intact
rock or along discontinuities. This test can measure peak and residual
strength as a function of stress normal to the shear plane.

The plane is selected so that is coincides and/or reflects a plane of


weakness in the rock, for example a joint, bedding plane or cleavage
Samples are tested after first subjecting them to confinement within
cement or other grout media and thereafter placing the confined sample
within a split jig assembly. A constant normal stress is applied to each
sample and then an increasing shearing load is applied, In each test, the
normal and failure shear stress conditions are recorded.

For each test specimen graphs of shear stress (or shear force) vs shear
displacement are plotted. Values of peak and residual strength are
extracted from the graph.

Version 1.10:Jan 2010: UNSW: GS 4-17


MINING GEOMECHANICS

250
a
P Peak strength
k 200
,
s
s
e
rt
s 150
r
a
e
h 100
S
50
Residual strength
0
0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0

Shear displacement, mm

Values of residual strength should only be interpreted if the sample is


sheared at constant normal stress and at least four consecutive
readings show  5% variation over a displacement of 1cm.

250
Peak strength
200
Shear stress, kPa

150 >5% fluctuation


 Cannot determine residual
100 strength

50

0
0.0 2.0 4.0 6.0 8.0 10.0 12.0 14.0 16.0
Shear displacement, mm

Multistage tests refer to the case when several tests are undertaken at
different normal stresses. The same discontinuity is often used for
multistage testing, however testing should only be at low normal
stresses as high normal stresses may cause shearing and result in
unreliable results. Due to the variability in results, the International
Society for Rock Mechanics suggests at least 5 such tests to gain a
result.

After plotting the peak shear stress with the appropriate normal stress,
the Coulomb failure locus for the rock defect can be determined. Even
with a large set of results as in the figure below, selecting the line of best
fit is a difficult task. In this case a line straight line of best fit has been
used as well as a comparison to Barton‟s criteria. However close
inspection may find a bilinear approach (according to Patton); or a
second (and more conservative) alternative would be to use the lower
bound values (the line connecting the lower most values).

Version 1.10:Jan 2010: UNSW: GS 4-18


MINING GEOMECHANICS

Many analytical and numerical methods in rock mechanics, however,


require the Coulomb parameters: that is, c and Φ. The mathematical
approach to exactly how a Barton curvilinear envelope is re-stated in
terms of linear Mohr Coulomb parameters is dealt with in more detail in
the following Topic, but at this stage it is possible to estimate the Mohr
Coulomb parameters by obtaining a tangent to the curve; that is;

  
 i  arctan  
 n 

and

ci   f   n tan i

to derive the straight line representative of the range of normal stress.


The subscript i in this case refers to „instantaneous‟ cohesion and friction
which would apply ONLY across the range of normal stress levels at
which the tangent to the envelope was drawn. If the range of normal
stress levels is σn1 to σn2 and the corresponding shear stress values are
τ1 and τ2, then;

  f 2  f 1 
i  arctan 
 and ci   f 1   n1 tan i
  n 2   n1 

Version 1.10:Jan 2010: UNSW: GS 4-19


Owner List
The Owners-SP 2657

Lot Unit Name and Main Contact Address Levy Address Notice Address Entitlement Balance
(-)prepaid
1 1 SAMIR BAHIG SHOUSHA 1 -1.46
2/62 MASCOT DRIVE, EASTLAKES NSW 2018 2/62 MASCOT DRIVE, EASTLAKES NSW 2018 2/62 MASCOT DRIVE, EASTLAKES NSW 2018
2 2 MR S B SHOUSHA 1 -2.45
2/62 MASCOT DRIVE, EASTLAKES NSW 2018 2/62 MASCOT DRIVE, EASTLAKES NSW 2018 2/62 MASCOT DRIVE, EASTLAKES NSW 2018
3 3 EDWARD ADHIKARI & DAISY BISWAS 1 0.00
3/62 MASCOT DRIVE, EASTLAKES NSW 2018 3/62 MASCOT DRIVE, EASTLAKES NSW 2018 3/62 MASCOT DRIVE, EASTLAKES NSW 2018
4 4 SEAN & CATHERINE O'DONOGHOE 1 0.00
4/62 MASCOT DRIVE, EASTLAKES NSW 2018 4/62 MASCOT DRIVE, EASTLAKES NSW 2018 4/62 MASCOT DRIVE, EASTLAKES NSW 2018
5 5 MR DAVID HENRY POPPLEWELL 1 0.00
5/62 MASCOT DRIVE, EASTLAKES NSW 2018 5/62 MASCOT DRIVE, EASTLAKES NSW 2018 5/62 MASCOT DRIVE, EASTLAKES NSW 2018
6 6 MR G & MRS U KASSAR 1 0.00
2 KYOGLE STREET, EASTLAKES NSW 2018 2 KYOGLE STREET, EASTLAKES NSW 2018 2 KYOGLE STREET, EASTLAKES NSW 2018
7 7 MR J & MRS M REPETTO 1 0.00
42 HIGHCLERE AVENUE, ROCKDALE NSW 42 HIGHCLERE AVENUE, ROCKDALE NSW 42 HIGHCLERE AVENUE, ROCKDALE NSW
2216 2216 2216
8 8 TRAM THI NGOG LE 1 0.00
8/62 MASCOT DRIVE, EASTLAKES NSW 2018 8/62 MASCOT DRIVE, EASTLAKES NSW 2018 8/62 MASCOT DRIVE, EASTLAKES NSW 2018
9 9 MR & MRS MAHMOUD 1 404.80
9/62 MASCOT DRIVE, EASTLAKES NSW 2018 9/62 MASCOT DRIVE, EASTLAKES NSW 2018 9/62 MASCOT DRIVE, EASTLAKES NSW 2018
10 10 YOHANES SOE & HANG LIN WIE 1 404.80
10/62 MASCOT DRIVE, EASTLAKES NSW 10/62 MASCOT DRIVE, EASTLAKES NSW 10/62 MASCOT DRIVE, EASTLAKES NSW
2018 2018 2018
11 11 MESSRS T, A & J & MRS K JAKUS 1 0.00
11/62 MASCOT DRIVE, EASTLAKES NSW 11/62 MASCOT DRIVE, EASTLAKES NSW 11/62 MASCOT DRIVE, EASTLAKES NSW
2018 2018 2018
12 12 MRS Y KASSAR 1 0.00
68 FLORENCE AVENUE, EASTLAKES NSW 68 FLORENCE AVENUE, EASTLAKES NSW 68 FLORENCE AVENUE, EASTLAKES NSW
2018 2018 2018

28/08/2009 18:49 Mark Gitman MG Strata and BMC Management Pty Ltd Page 1
Rock
MINING GEOMECHANICS

Mechanics

5
TOPIC 5: ROCK MASS
STRENGTH

5.1 Introduction
Previously we have examined the behaviour of intact rock and rock
defects (which separated masses of intact rock) from the perspective of
yield criteria. A rock material (intact rock PLUS discontinuities) thus
would usually behave quite differently from rock defects or intact
masses alone.

Consequently, rock mass behaviour will lie somewhere between these


two extremes of rock material and rock defect behaviour.

Version 1.10:Jan 2010: UNSW: GS 5-1


MINING GEOMECHANICS

Often we need to estimate the strength and deformation characteristics


for the assessment of overall rock mass stability in open pit slopes, or
the competency of the rock mass around underground excavations
These characteristics are associated with the large scale behaviour of
the rock mass and as such it is often not practical to establish these
properties from laboratory testing. In these situations we use empirical
relationships based on rock mass classifications to aid in these design
studies.

Determination of rock mass characteristics is part of the step of


engineering analysis. Following this, local and global excavation design
can be addressed after which monitoring will be used to determine the
reliability of the initial estimation.

Site Characterisation (Data collection and Analysis)

Mine model formulation (Engineering Analysis)

Design (Global & local)

Implementation

Rock Mass performance monitoring

Retrospective Analysis

5.1.1 Learning outcomes


By the time you have completed this topic you will be able to:

1. Evaluate examples of anisotropy of a single discontunity in a


rock mass.
2. Calculate the peak and post peak strength of a rock mass
based on an empirical criterion
3. Calculate the deformability characteristics of a rock mass
based on an empirical criterion
4. Convert from the Hoek and Brown empirical rock mass
strength model to the Coulomb equivalent instantaneous
cohesion and friction values.

Version 1.10:Jan 2010: UNSW: GS 5-2


MINING GEOMECHANICS

To assist you in completing this Topic, it is suggested you access


www.rocscience.com and register and download the free software
package RocLabv1.exe.

5.2 Anisotropic Strength Concepts


A rock mass can be defined as an aggregate of solid rock material
which contains pervasive and discrete structural features. Behaviour is
controlled by the geometrical nature and strength of the geological
discontinuities of the intact rock bridges. The rock mass strength is a
portion of the intact strength due to the presence of the geological
discontinuities.

 Intact
Rock

Rock Mass

Single discontinuity



As a first approach to examining this combined behaviour we can


examine the effect of joints on a rock mass strength, we can consider
anisotropic rock mass behaviour. Anisotropy can be;

 inherent - where planes of weakness occur associated with the


original rock mass; typical of shales, slates, phyllites, gneisses
and schists.
 induced - occurs after the rock is formed (e.g. joints, faults).

Only the simplest form of anisotropy is considered here (transverse


isotropy) because of computational complexity and the difficulty in

Version 1.10:Jan 2010: UNSW: GS 5-3


MINING GEOMECHANICS

determining the necessary elastic constants of the more complex


versions. The peak strengths developed by transversely isotropic rocks
in compression (uniaxial or triaxial) vary with the orientation of the plane
of isotropy, discontinuity plane or plane of weakness, with respect to the
principal stress directions, as shown below.

Uniaxial compressive strength can vary by a factor of 5 depending on


the direction of loading and properties are dependent on the strength
parallel to plane of weakness rather than normal to it.

Jaeger‟s analysis of rock containing well-defined, parallel planes of


weakness whose normals are inclined at an angle β to the major
principal stress direction as shown below. Each plane of weakness has
a limiting shear strength defined by Coulomb‟s criterion;

 w   n tan w  cw

Where;

the subscript w refers to plane of weakness strength parameters

σ1

σ3

Slip on the plane of weakness will become occur when the shear stress
on the plane is just greater than the shear strength, τw. Referring to

Version 1.10:Jan 2010: UNSW: GS 5-4


MINING GEOMECHANICS

Topic 3 for further development of the Coulomb yield criterion, but in this
case for a discontinuity or plane of weakness gives;

2cw   3 sin 2  tan w 1  cos 2 


 1w 
sin 2  tan w 1  cos 2  or

2cw   3 tan w 
 1w   3 
 tan w 
sin 2  1  
 tan  

Which gives a theoretical result similar to that shown below (LHS);

For values of β approaching 90° and in the range 0° to Φw°, slip on the
plane of weakness cannot occur, and the peak strength is governed by
some other mechanism, probably shear fracture through the rock
material in a direction not controlled by the plane of weakness.

The variation of peak strength with the angle β predicted by this theory is
in reasonable agreement with actual tests illustrated above (RHS),
although showing pronounced minima, do not follow quite the same
shape. This suggests that the two-strength model developed here
provides an oversimplified representation of strength variation in
anisotropic rocks. More complex models, not discussed here, are
available to overcome this simplification. One example is the adaptation
of the Hoek - Brown failure criterion introduced in Topic 3.

Version 1.10:Jan 2010: UNSW: GS 5-5


MINING GEOMECHANICS

5.3 Rock Mass Strength


The ideal failure criterion should:

 be compatible with the results of laboratory tests on intact rock


material;
 be able to deal with jointed rock masses; and
 be relatively straightforward to use in practical engineering.

The Hoek - Brown empirical rock mass failure criterion is the most
commonly used in practice. Details are presented in Topic 3, but for
review, the failure criterion for jointed rock masses is based on an
assessment of the interlocking rock blocks and the condition of the
surfaces between these blocks. The most general form of the Hoek-
Brown criterion for jointed rock masses is:
a
  
 1   3   ci  mb 3  s 
  ci 

Where

1 = major principal or maximum effective stress at failure;

3 = minor principal or minimum effective stress at failure;

 ci = UCS of the intact rock material;

mb = material constant for the rock mass.

s,a = constant, related to rock mass characteristics

Constants mb, s and a are difficult to calculate and are generally


estimated based on rock mass classification methods as discussed in
Topic 2.

For rock masses, as opposed to intact rock, the GSI can be used as an
alternate source for the variable mb where;
 GSI 100 
 
 2814 D 
mb  mi e

„D‟ refers to a disturbance factor in the range 0 to 1 which depends on


the amount of disturbance caused by blast damage and stress
relaxation.

Version 1.10:Jan 2010: UNSW: GS 5-6


MINING GEOMECHANICS

For Underground

 Excellent blasting / controlled blasting techniques: D = 0


 Poor blasting (with local damage 2 to 3 m into surrounding rock
mass) : D = 0.8

For Surface (In large scale open pits):

 Production blasting D = 1.0

Version 1.10:Jan 2010: UNSW: GS 5-7


MINING GEOMECHANICS

For the constant s; (rock masses);


 GSI 100 
 
9 3 D 
s  e

And exponent „a‟ in more general applications of the criterion to a rock


mass, the exponent (fixed at 0.5 in the original criterion for intact rock)
can be determined from;

  GSI   20 
   
e  15 
e  3 
a  0.5 
6

The uniaxial compressive strength of the rock mass is obtained by


setting σ3 = 0 in the Hoek and Brown equation, giving:

 c   ci s a

and, the tensile strength of the rock mass is:

s ci
t  
mb

For the uniaxial compressive strength (UCS, σc) of the rock mass given
above, failure initiates at the boundary of an excavation when UCS is
exceeded by the stress induced on that boundary. The failure
propagates from this initiation point into a biaxial stress field and it
eventually stabilizes when the local strength, defined Hoek and Brown
equation, is higher than the induced stresses around the excavation at
that point. Most numerical models can follow this process of fracture
propagation and this level of detailed analysis is very important when
considering the stability of excavations in rock and when designing
support systems.

There are applications when it is useful to consider the overall behaviour


of a rock mass rather than the detailed failure propagation process
described above. Pillar design would be an example where knowledge
of the overall strength of the pillar is required, in place of the extent of
fracture propagation in the pillar.

This leads to the concept of a global “rock mass strength” (σcm) and
Hoek and Brown proposed that this could be estimated for the stress
range t < 3 < 0.25.ci as:

Version 1.10:Jan 2010: UNSW: GS 5-8


MINING GEOMECHANICS

  mb  
a 1

 mb  4s  amb  8s   s 


 cm   ci   4  
 21  a 2  a  
 
 

Alternatively, rock mass compressive strength (σcm, MPa) can be


estimated from GSI, mi and UCS (intact rock values) following Marinos
and Hoek (2000) as;


 cm  0.0034 mi 0.8 c 1.029  0.025 exp 0.1m  i

GSI

5.3.1 Hoek-Brown and Coulomb equivalents


Many numerical models and limit equilibrium analyses are based on the
Mohr-Coulomb parameters, c and . This means couching the (non-
linear) Hoek-Brown criterion in terms of  and 'n (where n is the
(effective) normal stress across the failure surface through the rock
mass).

This is done by fitting an average linear relationship to the curve


generated by solving the Hoek and Brown equation for a range of minor
principal stress values defined by σt < σ3 <σ3max, as illustrated on the
next page. The fitting process involves balancing the areas above and
below the Mohr-Coulomb plot.

This results in the following equations for the angle of friction Φi and
cohesive strength ci:

 6amb s  mb 3n  
a 1
i  sin 1  a 1 
 21  a 2  a   6 amb s  m 
b 3n  

ci 
 1  2a s  1  a mb 3n s  mb 3n a 1 
ci

1  a 2  a  1  6amb s  mb 3n  


a 1

1  a 2  a 
Where;

 3 max
 3n 
 ci

The results of the studies for both deep and shallow tunnels are used to
derive an expression for σ3max such that;

Version 1.10:Jan 2010: UNSW: GS 5-9


MINING GEOMECHANICS

0.94
 
 3 max  0.47 cm  cm 
 h 

Where;

γ = unit weight of the rock mass

h = depth of the tunnel below surface. In cases where


the horizontal stress is higher than the vertical
stress, the horizontal stress value should be used.

This approach applies to all underground excavations, which are


surrounded by a zone of failure that do not extend to surface.

An example of the results, which are obtained from this analysis, is


given in the figures on the next pages from the RocLab software.

Version 1.10:Jan 2010: UNSW: GS 5-10


MINING GEOMECHANICS

Version 1.10:Jan 2010: UNSW: GS 5-11


MINING GEOMECHANICS

5.3.2 Hoek-Brown and Coulomb equivalents from GSI


An alternative approach to estimating Mohr-Coulomb shear strength
parameters from the Hoek and Brown equation for rock masses involves
the use of the rock mass GSI value, as discussed in Topic 2.

To determine c‟ and Φ’ , this approach required the uniaxial


compressive strength of intact rock (σci) and the Hoek and Brown
constant mi (also for intact rock). These values are then related to the
GSI rock mass classification value as described in Topic 2.

Plots of the values of the ratio c‘/ci and the friction angle ', for different
combinations of GSI and mi are created as design charts (below and
next page) for estimating the cohesion and friction angles.

Version 1.10:Jan 2010: UNSW: GS 5-12


MINING GEOMECHANICS

5.4 Rock Mass Deformability


The term deformability refers to the response of the material/ defect/
rock-mass to stress and includes an understanding of:

 the strains and displacements generated as stresses change


(constitutive behaviour);
 at what level failure occurs (strength); and
 the strains and displacements generated over time at more-or-
less constant stress levels (temporal or creep behaviour).

5.4.1 Background to deformability


A force-displacement curve can be obtained during the uniaxial
compression testing of a rock specimen. If the test is conducted with
strain as the independent variable (i.e. increasing the strain at a constant
rate) then under these conditions the complete force-displacement curve
can be obtained. This curve and it‟s interpretation forms the basis of
determining the parameters which describe rock strength;

 E, Young‟s Modulus

Version 1.10:Jan 2010: UNSW: GS 5-13


MINING GEOMECHANICS

  Poisson‟s ratio
 K, Bulk modulus
 G, Shear modulus

Fracture is the formation of planes of separation in the rock material. It


involves the breaking of bonds to form new surfaces. The onset of
fracture does not necessarily mean the rock has failed or peak strength
has been achieved.

Strength, or peak strength, is the maximum stress that the rock can
sustain under a given set of conditions. In the uniaxial compression test
illustrated on the next page, it corresponds to the maximum load-bearing
capacity of the rock, referred to as UCS (uniaxial compressive strength,
or σc.)

After peak strength has been exceeded, the rock may still have some
load-carrying capacity or strength. The minimum or residual strength is
reached generally only after considerable post-peak deformation. The
shape of this post-peak curve is related to the type of post-peak failure
that occurs;

 Brittle fracture is the process by which sudden loss of strength


occurs across a plane following little or no permanent (plastic)
deformation. It is usually associated with strain-softening or
strain-weakening behaviour of the rock.
 Ductile deformation occurs when the rock can sustain further
permanent deformation without losing load-carrying capacity

Version 1.10:Jan 2010: UNSW: GS 5-14


MINING GEOMECHANICS

Semi-brittle fracture is a combination of the above two post-peak failure


modes. The shape of the post-peak curve will depend also on how the
test is programmed. The curve can be considered in two main sections:

 The pre-peak and peak


 The post peak

The engineering aspects that we are interested in are :

 The peak strength


 The modulus of elasticity (Young‟s Modulus E)

5.4.2 Young’s Modulus


Young‟s modulus is derived from the gradient of the slope of the linear
section of the stress-deformation curve. In reality, the slope is often
constantly changing, therefore three methods are considered.

Tangent Modulus
Et is the slope at some fixed percentage, generally 50% of the peak
strength σc. (designated here by σc./2).

Referring to the results of a uniaxial test shown below;

Version 1.10:Jan 2010: UNSW: GS 5-15


MINING GEOMECHANICS

110
σc = 104MPa
100
90
80 Slope = Es

70
Stress (MPa)
60
50 σc/2 = 52MPa

40
30
20
10 Slope = Et, Eav
0
0 0.001 0.002 0.003 0.004
Strain (dimensionless )

52
Et   44GPa
0.00168  0.0005 

Average Young’s Modulus


Eav is the average slope of the more-or-less straight portion of the axial
stress-strain curve. Using the above case; Eav = Et = 44GPa

Secant Young’s Modulus


Esecant is the slope of a straight line joining the origin of the axial stress-
strain curve to a point on the curve at some fixed percentage of the peak
strength. In this example the secant modulus at peak strength is;

104
Et   32.5GPa
0.0032 

5.4.3 Pre-peak and peak behaviour


The typical rock response to applied stress generally appears in the
form of inelastic or non-linear stress/strain behaviour, as shown:

Version 1.10:Jan 2010: UNSW: GS 5-16


MINING GEOMECHANICS

At the very beginning, up to the limit of region A, the curve has an


initial portion that is concave upwards. This occurs for two reasons:

1. The lack of perfect specimen preparation - the ends of


the cylinder are not parallel
2. The closing of microcracks within the rock material
In the second case, existing cracks within the rock will “seal up” under
the action of applied external stress, resulting in formation of an inelastic
or curved stress/strain envelope. This behaviour will normally occur in
rock specimens which have been destressed (i.e.- by the process of
diamond drilling recovery) and then reloaded.

Region B represents the linear portion of the curve. The rock behaviour
is more or less analogous to the ideal „elastic‟ rock. The stress/strain
behaviour follows an ideal elastic path. This condition is the general one
for confined rock materials, existing at depth, which have not undergone
destressing.

Region C indicates the start of yield and represents the “theoretical


point” at which the material stops behaving in an elastic manner and the
deformation becomes permanent, or plastic. The rock specimen will
undergo formations of new microcracks due to internal fracturing. The
microcracks coalesce into macroscopic planes which ultimately lead to
failure. Upon being subjected to a stress level equal to the unconfined
compressive strength of the rock, sample shear failure will result. The
formation of microcracks and macroscopic shear planes will produce
inelastic stress/strain behaviour, which is typical for rock which is
unconfined or confined but acted upon by high stress.

Microcracking generates noises or acoustic emissions that can be


monitored. Significant acoustic emission is found to occur starting at a

Version 1.10:Jan 2010: UNSW: GS 5-17


MINING GEOMECHANICS

level around 50% of c. and continues until complete failure. This will be
considered in a later Topic.

5.4.4 Post-peak behaviour


In other forms of engineering, e.g. the strength of concrete in structural
engineering, if the applied stress reaches the compressive strength,
there can be catastrophic consequences. Rocks on the other hand can
carry load beyond what would normally be considered failure. This is
residual strength referred to in the diagram below. The value of residual
strength depends on the type of post-peak behaviour

100

90

80

70
Stress (MPa)

60

50 Brittle post-peak residual


Ductile post-peak residual
40

30

20

10

0
0 0.001 0.002 0.003 0.004
Strain (dimensionless )

The rock material behaves in a brittle manner post-peak when strain


softening occurs. That is, there is a significant loss in the ability of the
rock material to carry load.

The rock material behaves in a ductile manner post peak when strain
hardening occurs. That is, the rock material continues to carry an
increased load albeit at a lower modulus.

However, remember that the post-peak (residual) behaviour is also


associated with the type of testing machine used.

Version 1.10:Jan 2010: UNSW: GS 5-18


MINING GEOMECHANICS

5.4.5 Site-specific rock mass modulus


Generally, it is impractical to determine the site-specific rock mass
modulus. Empirical estimation methods are again generally used.
From Topic 2 the following formulae for modulus of deformability (GPa)
were introduced;

 Barton
Emass  25log10 Q for Q>1 and generally hard rocks

 Serrafim and Pereira, where UCS < 100MPa;


 GSI 10 
c  
E mass  10  40 
100

 Serrafim and Pereira, where UCS > 100MPa


 GSI 10 
 
E mass  10  40 

 Barton
1
 Q c  3
E mass  10  where c = UCS of intact rock mass
 100 

 15 log10 Q  40 
 
E mass  10  40 

Further modification were made to include the value of the Disturbance


factor (D) such that:

 where UCS < 100MPa;


 GSI 10 
 D   c  
E mass  1   10 40 

 2  100

 where UCS > 100MPa


 GSI 10 
 D  
E mass  1  10  40 

 2

Version 1.10:Jan 2010: UNSW: GS 5-19


MINING GEOMECHANICS

The graph below (from Hoek, 2004) illustrates the differences between
several of the relationships above:

5.4.1 Poisson’s Ratio


It may also be necessary to estimate a Poisson‟s Ratio of the rock
mass. This will be dependent on the intact rock properties and the
frequency and characteristics of the defects within the rock mass. In
practice only a very rough estimate can be made and would typically
vary from 0.2 for very good quality rock masses to 0.3 for very poor
quality rock masses.

5.5 Summary
In summary, the Hoek-Brown failure criterion should only be applied to
those rock masses in which there are a sufficient number of closely
spaced discontinuities, with similar surface characteristics, such that
isotropic behaviour involving failure on discontinuities can be assumed.

Version 1.10:Jan 2010: UNSW: GS 5-20


MINING GEOMECHANICS

When the structure being analysed is large and the block size small in
comparison, the rock mass can be treated as a Hoek-Brown material.

The figure below summarises the applications as discussed here and in


previous Topics.

APPLICABILITY

Use INTACT ROCK mi


and Hoek -Brown
equations

Hoek -Brown not


applicable. Use
Coulomb criterion

Hoek -Brown not


applicable. Use
Coulomb
ANISOTROPIC
analysis

Use HOEK -BROWN


mb , a and s values f or
ROCK MASSES with
>4 joint sets of unif orm
strength

Use HOEK -BROWN


mb , a and s values f or
ROCK MASSES

References
Hoek, E 1983, Strength of Jointed rock Masses, Geotechnique, 33, 3, pp187-223
Hoek E. & Brown, E.T. 1997 Practical Estimates of Rock Mass Strength, Int. J.
Rock Mech. Min. Sci., Vol 34, No 8, pp1165-1186
Hoek, E. Rock Engineering. Course Notes. Available via www.rocscience.com
Marinos, P and Hoek, E. 2000. GSI – A geologically friendly tool for rock mass strength
estimation. Proc. GeoEng2000 Conference, Melbourne.

Version 1.10:Jan 2010: UNSW: GS 5-21


Owner List
The Owners-SP 2657

Lot Unit Name and Main Contact Address Levy Address Notice Address Entitlement Balance
(-)prepaid
1 1 SAMIR BAHIG SHOUSHA 1 -1.46
2/62 MASCOT DRIVE, EASTLAKES NSW 2018 2/62 MASCOT DRIVE, EASTLAKES NSW 2018 2/62 MASCOT DRIVE, EASTLAKES NSW 2018
2 2 MR S B SHOUSHA 1 -2.45
2/62 MASCOT DRIVE, EASTLAKES NSW 2018 2/62 MASCOT DRIVE, EASTLAKES NSW 2018 2/62 MASCOT DRIVE, EASTLAKES NSW 2018
3 3 EDWARD ADHIKARI & DAISY BISWAS 1 0.00
3/62 MASCOT DRIVE, EASTLAKES NSW 2018 3/62 MASCOT DRIVE, EASTLAKES NSW 2018 3/62 MASCOT DRIVE, EASTLAKES NSW 2018
4 4 SEAN & CATHERINE O'DONOGHOE 1 0.00
4/62 MASCOT DRIVE, EASTLAKES NSW 2018 4/62 MASCOT DRIVE, EASTLAKES NSW 2018 4/62 MASCOT DRIVE, EASTLAKES NSW 2018
5 5 MR DAVID HENRY POPPLEWELL 1 0.00
5/62 MASCOT DRIVE, EASTLAKES NSW 2018 5/62 MASCOT DRIVE, EASTLAKES NSW 2018 5/62 MASCOT DRIVE, EASTLAKES NSW 2018
6 6 MR G & MRS U KASSAR 1 0.00
2 KYOGLE STREET, EASTLAKES NSW 2018 2 KYOGLE STREET, EASTLAKES NSW 2018 2 KYOGLE STREET, EASTLAKES NSW 2018
7 7 MR J & MRS M REPETTO 1 0.00
42 HIGHCLERE AVENUE, ROCKDALE NSW 42 HIGHCLERE AVENUE, ROCKDALE NSW 42 HIGHCLERE AVENUE, ROCKDALE NSW
2216 2216 2216
8 8 TRAM THI NGOG LE 1 0.00
8/62 MASCOT DRIVE, EASTLAKES NSW 2018 8/62 MASCOT DRIVE, EASTLAKES NSW 2018 8/62 MASCOT DRIVE, EASTLAKES NSW 2018
9 9 MR & MRS MAHMOUD 1 404.80
9/62 MASCOT DRIVE, EASTLAKES NSW 2018 9/62 MASCOT DRIVE, EASTLAKES NSW 2018 9/62 MASCOT DRIVE, EASTLAKES NSW 2018
10 10 YOHANES SOE & HANG LIN WIE 1 404.80
10/62 MASCOT DRIVE, EASTLAKES NSW 10/62 MASCOT DRIVE, EASTLAKES NSW 10/62 MASCOT DRIVE, EASTLAKES NSW
2018 2018 2018
11 11 MESSRS T, A & J & MRS K JAKUS 1 0.00
11/62 MASCOT DRIVE, EASTLAKES NSW 11/62 MASCOT DRIVE, EASTLAKES NSW 11/62 MASCOT DRIVE, EASTLAKES NSW
2018 2018 2018
12 12 MRS Y KASSAR 1 0.00
68 FLORENCE AVENUE, EASTLAKES NSW 68 FLORENCE AVENUE, EASTLAKES NSW 68 FLORENCE AVENUE, EASTLAKES NSW
2018 2018 2018

28/08/2009 18:49 Mark Gitman MG Strata and BMC Management Pty Ltd Page 1
Rock
MINING GEOMECHANICS

Mechanics

6
TOPIC 6: ROCK
TESTING

6.1 Introduction
When considering the design excavations in rock masses, it is important
to;

 Have a sound understanding of engineering properties of the


materials being dealt with;
 Have an understanding what tests should be done in order to derive
the required properties;
 Understand the potential errors that may occur in the testing
methods AND the limits of accuracy and reliability.

For a rock mechanics design, you not only need to be able to select the
required test for the required property, but you also need the knowledge
to question whether the result makes sense. To fulfil these outcomes, a
sound understanding of engineering properties and how these are
determined, is required.

As Topic 6 is essentially an extension of Topics 3, 4, and 5, in which


some of these tests and results have already been introduced, it takes
its place in the first level of the general rock engineering process as
shown on the next page.

CRITICALLY, when testing rock to obtain data for use in design, is that
the boundary conditions applied to the test specimen should simulate
those imposed on the rock in situ. Practically, this is difficult to achieve
and general practice is to study the behaviour of the rock under known
uniform applied stress systems.

Version 1.10:Jan 2010: UNSW: GS 6-1


MINING GEOMECHANICS

Site Characterisation (Data collection and Analysis)

Mine model formulation (Engineering Analysis)

Design (Global & local)

Implementation

Rock Mass performance monitoring

Retrospective Analysis

6.1.1 Learning Objectives


By the time you have completed this topic you will be able to:

1. Describe the various tests used to determine the engineering


properties of rocks
2. Determine the engineering properties of rock from given data
3. Evaluate the shortcomings of each of the tests as indicators of
the large scale properties of rock.
4. Select the required test for the required property
5. Make engineering observations from the results of tests

6.2 Engineering Tests and Properties


The table on the next page comprises some of the more and less
common tests that will be used to determine engineering properties.
The most common tests are the;

 Uniaxial Compressive Strength test, commonly known as the UCS


 Uniaxial deformability tests, to determine Young‟s modulus, and
Poisson‟s ratio.

Version 1.10:Jan 2010: UNSW: GS 6-2


MINING GEOMECHANICS

 Triaxial tests, and multistage triaxial tests were traditionally not


commonplace in hard rock mining, with both normally reserved to soil
mechanics. However they have become increasingly used in the
area of backfill testing.
 The swelling pressure and slake durability tests are used to assess
the resistance offered by a rock sample to weakening and
disintegration when subjected to cycles of drying and wetting used in
coal mining.
 Shear box testing - used to gauge shear strength.

Less common tests are the Brazilian Tensile Test and the Creep Tests.
Both of these are generally reserved for the domain of research.

More Common Tests Property


Uniaxial Compressive Test (Uniaxial compressive) Peak
strength, Young‟s Modulus,
Poisson‟s Ratio
Triaxial Compressive Test Strength envelope (from multiple
samples) and Mohr Coulomb
parameters
Multistage Triaxial test Strength envelope (from a single
sample) and Coulomb
parameters
Shear box test Shear strength
Point Load Index Strength
Dynamic elastic Tests Dynamic elastic constants
Slake Durability Index Swelling strain and slake
durability
Less Common Tests Property
Brazilian Tensile Test Tensile strength
Creep Test Time-dependent behaviour

There are even more tests than are presented in this list. However, this
Topic will focus on the tests that determine the following properties :

 Strength and strength indices


 Deformability
 Slaking

Shear strength tests have been dealt with previously and dynamic
elastic constants and associated tests will be introduced in a later Topic.

Version 1.10:Jan 2010: UNSW: GS 6-3


MINING GEOMECHANICS

6.2.1 Testing regime


Rock properties can be determined both from laboratory and field tests.
In the laboratory we can derive material/engineering properties of rocks
and discontinuities under so called „Controlled‟ conditions, with testing
procedures adhering to the International Society of Rock Mechanics
(ISRM), ISO 9001, and Australian Standards.

In field, we can derive properties from Index tests, which give us a


comparative measure of, for example, strength. Field tests must also be
conducted in accordance with the relevant standards, however the
nature of the tests means that they are less controlled and therefore less
reliable. They should be considered specific only for the mining
operation where they were conducted.

Rock elastic and inelastic behaviour are influenced by the test


parameters themselves; such as size, shape of sample; system
stiffness, rate of loading, etc. These influences apply to all testing and
underline the point that all properties (test results) are a function of the
rock materials and the testing conditions (size, shape, system stiffness,
etc), and not an inherent property. It is important to remember this when
examining test result data. Always remember to ask

 Why you are requesting tests, and what their purpose is?
 What are the test conditions versus the in situ conditions?
 Will you be able to extrapolate between the two conditions – sample
testing volume and rock (mass) volume?

6.3 Strength Tests


Strength is probably the first property that a mining engineer will
consider when designing an excavation. We can use both laboratory
and field tests to classify the rock strength, including:

 Unconfined compressive strength (UCS) test (lab test)


 Confined compressive strength (Triaxial) test (lab test)
 Tensile strength
 Direct Tensile Test (lab test)
 Indirect Tensile (Brazilian) Test (lab test)
 Point Load (Index) Strength (field test)

Version 1.10:Jan 2010: UNSW: GS 6-4


MINING GEOMECHANICS

6.3.1 Unconfined Compressive Strength (UCS)


The uniaxial or unconfined compressive strength is the property most
often quoted to characterise the mechanical behaviour of rock. This
parameter is the strength exhibited when a rock sample is subjected to
compression stresses in a uniaxial fashion. Typically, a cylindrical
specimen of rock, or core, is recovered by drilling into the rock matrix.
Such core materials are loaded axially, in compression, with no radial
(σ3) stresses being present during failure testing to provide lateral
confinement.

Standard test procedure and interpretation


Suggested techniques for determining the uniaxial compressive strength
and deformability of rock material are given by the International Society
for Rock Mechanics Commission on Standardization of Laboratory and
Field Tests (ISRM Commission, 1979). The essential features of the
recommended procedure are:

1. The test specimens should be right circular cylinders having a


height to diameter ratio of 2.5–3.0 and a diameter preferably
of not less than NX core size, approximately 54 mm. The
specimen diameter should be at least 10 times the size of the
largest grain in the rock.
2. The ends of the specimen should be flat to within 0.02 mm
and should not depart from perpendicularity to the axis of the
specimen by more than 0.001 rad or 0.05 mm in 50 mm.
3. The use of capping materials or end surface treatments other
than machining is not permitted.
4. Specimens should be stored, for no longer than 30 days, in
such a way as to preserve the natural water content, as far as
possible, and tested in that condition.
5. Load should be applied to the specimen at a constant stress
rate of 0.5–1.0 MPa s−1.
6. Axial load and axial and radial or circumferential strains or
deformations should be recorded throughout each test.
7. There should be at least five replications of each test.

The following Figure shows an example of the results obtained. Note


that;

 the axial force has been divided by the initial cross-sectional area of
the specimen to give the average axial stress, σa,
 axial strain, εa, is determined from the ratio of initial specimen length
to loaded length

Version 1.10:Jan 2010: UNSW: GS 6-5


MINING GEOMECHANICS

 radial strain, εr is determined from the ratio of initial specimen radius


(at centre of specimen) to loaded radius.
 An alternative is to use strain gauges mounted on the specimen
(see later illustration under triaxial testing) to measure these
values directly.

Where post-peak deformations are recorded, the cross-sectional area


may change considerably as the specimen progressively breaks up. In
this case, it is preferable to present the experimental data as force–
displacement curves.

In terms of progressive fracture development and the accumulation of


deformation, the stress-strain or load-deformation responses of rock
material in uniaxial compression generally exhibit the four stages
discussed in Topic 5;

1. An initial bedding down and crack closure stage


2. The stage of elastic deformation until an axial stress of σci is
reached at which stable crack propagation is initiated.
3. When axial stress reaches σcd, unstable crack growth and
irrecoverable deformations begin.
4. Peak or uniaxial compressive strength, σc, or UCS is reached,
following which post-peak behaviour is initiated. Unconfined
sample failure can occur in several modes including a process
of sliding on shear planes at some oblique angle to the axial
stress.
However, not all rock types follow this process and compressive
strength tests will generally fall into one of three mode types:

Version 1.10:Jan 2010: UNSW: GS 6-6


MINING GEOMECHANICS

 Type A (straight line) indicates a linear elastic behaviour indicating a


constant value of the modulus of elasticity till the point of failure. This
type of curve is exhibited by most of the igneous rocks, such as
basalt, diabase, gabbro and other very strong rocks such as
quartzite, very strong sandstone, limestones, durain (coal) etc
 Type B (convex towards the stress axis) indicates a relationship
where there is pronounced strain with every increment of load. The
value of the modulus of elasticity is highest in the early stages of
loading, but continuously decreases. Such a behaviour is exhibited
by softer rocks such as shales, siltstones, tuff, softer limestones, well
cleated and bedded coal when tested parallel to the bedding planes,
etc. Typical strain-softening.
 Type C (convex towards the strain axis) indicates a stress-strain
curve where there is decreased strain with every increment in load.
The value of the modulus of elasticity is the lowest at the start of
loading but continuously increased. This behaviour is exhibited by
sandstones, coals and other rocks when loaded perpendicular to the
bedding planes, rock salt and certain metamorphic rocks. Typical
strain-hardening.

As shown above, the axial Young‟s modulus of the specimen varies


throughout the loading history and so is not a uniquely determined
constant for the material.

Failure modes
Several modes of failure experienced during testing:
 Shear failure is observed when single or multiple crack surfaces
propagate completely through the sample at some measured, and

Version 1.10:Jan 2010: UNSW: GS 6-7


MINING GEOMECHANICS

generally fixed, angle to the direction of axial stress application (LHS


illustration). Sliding of the individual, intact portions of the sample will
occur along the fracture surfaces which are created. The multiple
planes can occur in conjugate pairs (RHS illustration) where the
major angle is normal to the direction of least stress.

This will often preferentially occur along discontinuities such as


bedding planes if these exist at some oblique angle to the
direction of load.

 Axial splitting is observed when cracks parallel to the direction of the


axial load occur, suggesting that the bonds between the grains failed
in tension. The same failure mode may also be seen in underground
mines on hard rock pillars or of drive walls where it is also termed
„spalling‟ or fretting. The larger scale cases may also involve
bedding planes running parallel to the wall.

 Multiple cracking is observed when the rock disintegrates along


many planes of random direction. It is often a combination of the first
two failure types.

Version 1.10:Jan 2010: UNSW: GS 6-8


MINING GEOMECHANICS

6.3.2 Confined (Triaxial) Compressive Strength


This parameter is the sample strength exhibited when a rock sample is
subjected to compression stress when simultaneously held under radial
confinement (called “triaxial” confinement). Core samples are loaded
axially, in compression, between the sample end faces while lateral
(radial) compressive stress is also applied. The radial stresses which
are developed in order to confine the sample act such that (2 = 3),
thereby assuming that, at failure, the intermediate and minor principal
stress components acting upon specimens will be equal in magnitude.

Failure under triaxial confinement occurs at some level of applied axial


stress (1) which will always be greater in magnitude than the
unconfined compressive strength of the rock material. Typically,
triaxially-confined sample failure tests are conducted in triaxial cells,
where radial confinement is applied to rock core specimens by a
surrounding pressurized fluid bath of hydraulic oil (see figure below)

A general state of three-dimensional stress at a point can be


represented by three principal stresses, σ1, σ2 and σ3, acting on
mutually orthogonal planes. No shear stresses act on these planes. A
plane of particular interest is the boundary of an underground
excavation which is a principal plane except in the unusual case in
which a shear stress is applied to the boundary surface by the support.
The rock surrounding an underground excavation is rarely in a state of

Version 1.10:Jan 2010: UNSW: GS 6-9


MINING GEOMECHANICS

uniaxial compression. In the general case, away from the excavation


boundary or on the boundary when a normal support stress, σ3, is
applied, there will be a state of polyaxial stress (σ1 ≠ σ2 ≠ σ). The special
case in which σ2 = σ3 is called triaxial stress. It is this form of multiaxial
stress that is most commonly used in laboratory testing. On the
boundary of an unsupported excavation, σ3 = 0, and a state of biaxial
stress exists. Biaxial stress tests are very complex tests to perform and
can be replicated to an extent by triaxial tests.

Typical triaxial test results

Confining pressures (σ3)

The Figure above shows results from eight triaxial tests on the same
rock type, from 0 to 48.3MPa confining pressure. A number of important
features of the behaviour seen are that, with increasing confining
pressure;

1. the peak strength increases;


2. there is a transition from brittle to ductile behaviour;
3. the region incorporating the peak of the curve flattens and
widens;
4. the post-peak drop in stress to the residual strength reduces
and disappears at high values of confining pressure.
The confining pressure at which the post-peak reduction in strength
disappears and the behaviour becomes fully ductile (σ3 = 48.3 MPa in
the above Figure), is known as the brittle–ductile transition pressure and

Version 1.10:Jan 2010: UNSW: GS 6-10


MINING GEOMECHANICS

varies with rock type. In general, the more siliceous igneous and
metamorphic rocks such as granite and quartzite remain brittle at room
temperature at confining pressures of up to 1000 MPa or more. In these
cases, ductile behaviour will not be of concern in practical mining
problems.

6.3.3 Point Load Index (Is) Strength


The Point Load Test is a fast and convenient way to determine the
strength of a rock. Point load index tests are performed upon
unprepared or uncut lengths of diamond drill core which must exhibit a
minimum length-to-diameter aspect ratio of 1.5 :1 or greater. Because
point load test core samples do not require costly end preparation by
machining, such point load tests can be conducted using rudimentary
equipment, often in-situ, and during the early core recovery stages of
mine development.

The Point Load Strength test is intended as an index test for the strength
classification of rock materials. It may also be used to predict other
strength parameters with which it is correlated, for example the
unconfined compressive and the tensile strength. It is measured in
accordance with the procedures recommended in AS4133.4.1 usually
with NX-size core samples. The testing machine consists of a loading
frame, which measures the force required to break the sample, and a
system for measuring the distance between the two platen contact
points. Rock specimens in the form of either core, cut blocks, or irregular
lumps are broken by application of concentrated load through a pair of
spherically truncated, conical platens.

Version 1.10:Jan 2010: UNSW: GS 6-11


MINING GEOMECHANICS

In point load testing, core samples are subjected to diametral loading


(loading across the core diameter) between two conical indenters. Each
indenter is required to have a 600 conical-shaped tip, with the tip having
a 5 mm radius. Loads are applied between the conical indenters, across
the core diameter, until sample failure occurs, often by formation of a
single, planar tension crack.

In the following formulas, the Point Load Index strength (Is) is defined
as;

Diametral test

P
Is  2
De

Where;

P = measured value of force at which failure occurs (N) and De =


diameter of the core (m)

Axial, block and lump tests


0.5
 4A 
De   
  

Where A = the minimum cross sectional area of a plane through the


specimen and the platen contact points (D x W).

Version 1.10:Jan 2010: UNSW: GS 6-12


MINING GEOMECHANICS

A standardized strength value (for D = 0.05m), when a test is


performed at diameter De, (or 4A/π) is given as;
0.45
 De 
I s ( 50)  I s  
 0.050 


The standardised strength value can be used to ESTIMATE UCS from
the following formula;

 c  22 to 24I s (50)

Alternatively, use the data below to establish coefficient K values such


that;

c  k Is

Core diameter K
20mm 17.5
30mm 19
40mm 21
50mm 23
54mm 24
60mm 24.5

Note that the strength based on Point Load Index should only be stated
to nearest 5MPa.

6.3.4 Schmidt Rebound Hammer


This test was originally designed to measure joint wall compressive
strength (and was discussed in Topic 4). The Schmidt Hammer is a
portable device, by which a spring driven cylindrical hammer rebounds
off the rock surface; the rebound distance is considered to be a measure
of the rock quality. The hammer can be used directly on a rock surface,
or on a rock core. When an in situ block is tested and the block size is
large, the Schmidt hammer does measure the properties of the intact
rock; when the rock is fragmented, the measure will be of rock mass
quality rather than rock material.

The condition of the tested surface will have a significant effect on the
result, due to geometrical irregularities, or due to weathering of surface.
Thus the hammer should be used in various points in the immediate
vicinity of the measurement location. If surface irregularities are causing

Version 1.10:Jan 2010: UNSW: GS 6-13


MINING GEOMECHANICS

variation, discard the low values. If surface is weathered , then all


values are of significance.

Empirically determined curves show a relationship between Schmidt


Hardness and rock UCS depending upon rock density and the
orientation of the hammer relative to the surface relating the Schmidt
hammer readings. The vertical axis on the left hand side is UCS in
MPa; and on the RHS is the unit weight of the rock in kN/m3. The
horizontal axis (lower) represents the rebound distance for various
angles of the hammer 0, 45, 90, 135, and 180° towards the rock
surface.

Version 1.10:Jan 2010: UNSW: GS 6-14


MINING GEOMECHANICS

6.4 Tensile Strength


Although some rock failure criteria are based on tensile strength, mining
operations will rarely request tests for the property. Thus tensile testing
is generally confined to academic and research pursuits.

Tensile strength can be determined via direct or indirect methods. The


most common indirect tensile is the Brazilian Tensile Test.

Direct Tensile Test (σt)


The direct tensile test occurs when a core sample of rock is stressed, in
tension, along the sample axial direction. This is accomplished by
gripping the sample ends and subjecting the sample to an axial pulling
force. Failure occurs by propagation of cracks perpendicular to the
direction of the applied pulling force, or perpendicular to the specimen
long axis, as shown in the accompanying sketch.

The unconfined tensile strength (σt) represents the maximum tensile


stress that a sample can be subjected to before failing in tension. In
general, the magnitude of the unconfined compressive strength (σc) is
much greater than the magnitude of the unconfined tensile strength (σt).

Brazilian Tensile Tests, (σt)


Brazilian tensile strength test is performed by point loading thin discs of
core in compression between steel loading platens until the samples fail.
When failure occurs, diametral cracks form between the contact platen
points and across the single wafer diameter that is subjected to loading
forces. By subjecting the wafer sample to compressive force, it is
caused to fail in tension as though a tensile force had been exerted to
pull the wafer halves apart, as illustrated in the following sketch.

Version 1.10:Jan 2010: UNSW: GS 6-15


MINING GEOMECHANICS

The failure tensile stress applied is equivalent to the tensile strength (σt)
of the rock material. If the applied force (P) is measured in units of N
and test disc diameter (D) and thickness (t) in mm, then σt has units of
MPa.

0.636P
t 
D.t

6.5 Swelling and Slake-Durability Index


Properties
Rock properties, particularly at and near free surfaces, can change by
exfoliation, hydration, weathering, decrepitation (slaking), solution,
oxidation, abrasion, fatigue, and other chemico-hydro-thermo-
mechanical processes. In some shales and some volcanic rocks,
radical deterioration in rock quality occurs rapidly after a new surface is
uncovered.

Slake durability test is an index test for rock durability. The apparatus
consists of a drum of 140mm diameter and 100mm length with sieve
mesh forming the cylindrical walls (2mm opening). A 500g sample of
rock is broken into 10 lumps and loaded inside the drum, which is turned

Version 1.10:Jan 2010: UNSW: GS 6-16


MINING GEOMECHANICS

at 20 revolutions per minute in a water bath. After 10 minutes (1st cycle),


and 20 minutes (2nd cycle) of this slow rotation, the percentages of rock
retained inside the drum (for each cycle and on a dry weight basis), are
reported as the slake durability index, Id. The durability of the rock tested
is then classified according to the values of Id for each test cycle (Id1 and
Id2).

Gamble proposed the following durability classification:

Very high durability: Id1>99 and Id2>98.

High durability: 99> Id1>98 and 98> Id2>95.

Medium high durability: 98> Id1>95 and 95> Id2>85.

Medium durability: 95> Id1>85 and 85> Id2>60.

Low durability: 85> Id1>60 and 60> Id2>30.

Very low durability: Id1<60 and Id2<30.

6.6 Laboratory Testing Errors


Rock strength tests, which have been described in the previous
sections, are utilized to determine physical rock properties, and in
general, laboratory strength test analyses yield strength estimates for
rock samples that are higher than for similar materials in-situ due to
scale effects, the effects of rate of loading, sample shape effects such
as length-to-diameter ratio, and to conditions of sample confinement. In
all cases, data is usually obtained from a limited number of tests due to
the high cost of obtaining samples, preparing samples and conducting
tests. We will briefly look at seven of the aforementioned parameters
that may effect the results of UCS tests:

 The size (or scale) of the specimen


 Shape effect
 Uneven axial stress distribution
 Non-parallel end loading effects
 The specimen loading conditions – including confinement pressure
and loading rate (fast or slow)
 Environmental effects
 Sampling

Version 1.10:Jan 2010: UNSW: GS 6-17


MINING GEOMECHANICS

6.6.1 Size or Scale Effects of Samples


As sample size increases, the probability that each sample will include
or be influenced by the effects of fractures will increase. Core
specimens, recovered at much smaller size, may not intercept any joints
and the core strength will therefore not be affected by their presence.

Upon strength testing, such unjointed core samples will exhibit much
higher strength capabilities than will specimens which are much larger
and which will include joints or fractures (zones of weakness). In
general, as the size or volume of samples to be tested increases, the
strength of samples will exhibit a steady decrease, as indicated by the
diagram and relationship below. In this case, the relationship uses a
base case of a 50mm diameter core and is used to estimate the UCS
(σcd)of a larger sample (diameter d (mm)) such that;

 c d  50  0.18
 
 c 50  d 

6.6.2 Shape Effect of Sample


A cylindrically-shaped sample is best suited for strength testing of rock
materials because it is easy to create core samples in-situ by diamond
drilling, and the regular symmetry of samples assists in understanding
behaviour during testing..

Variation in the length-to-diameter ratio of core specimens is known to


affect the strength capabilities of samples. Strengths will generally
decrease as the sample length-to-diameter (L:D) ratio increases.

Version 1.10:Jan 2010: UNSW: GS 6-18


MINING GEOMECHANICS

 At very large (L:D) ratios, generally exceeding 5:1, samples become


subject to bending moment reactions and will fail in tension rather
than by shear fracture and sliding. They will show unrealistically low
compressive strengths.
 At very low (L:D) ratios, the physical geometry of the specimens will
be such that no shear failure surface will be able to propagate
completely through the specimens. Because shear sliding cannot
physically develop, shear failure is prevented from occurring and the
“short” samples will exhibit unrealistically high compressive
strengths.

The ISRM and Australian standards therefore recommended, that test


sample sizes always be prepared to aspects between 2/1 and 2.5/1 in
order to minimize the deleterious effects caused by under- or oversizing
of specimens.

6.6.3 Uneven Axial Stress Distribution


Rocks imperfections and asymmetric ends create some measure of
stress profile variation at the sample ends (loading positions) and
throughout the sample length.

Across the ends this usually results in high side stresses and lower
stresses developed through the mid-spans. Across the centre line of
each specimen (in the horizontal plane) low stress conditions exist at the
sample edges and higher axial stresses within the centre. The typical
horizontal plane stress development is illustrated on the next page.

Version 1.10:Jan 2010: UNSW: GS 6-19


MINING GEOMECHANICS

Irregularities can also be created at the ends of samples due to the


effects of confinement, or “clamping”, of the rock material between the
rock specimens and the metal loading plates (platens). The platens and
rock sample materials exhibit considerable mismatch in elastic moduli,
such that (Emetal >> Erock).

When an axial load is applied, both the rock and the platens will deform
axially and laterally in accordance to the Poisson ratio. The mismatch in
elastic moduli, means that the rock be able to deform to a much grater
degree than the metal platens. The less deformable platens will act to
resist lateral rock strains, creating a shear resistance at the rock/metal
interface and “clamping” the specimen ends rigidly.

The shear resistance will in turn induce a measure of radially-acting,


triaxial confining stress about the sample ends. The triaxial confinement
conditions which exist about sample ends create zones of triaxial
disturbance which extend up to one sample diameter into the samples

Version 1.10:Jan 2010: UNSW: GS 6-20


MINING GEOMECHANICS

from each plane of contact between the rock and steel end platens. The
triaxial confinement increases the effective strength the sample ends, so
samples will tend to fail at a higher indicated stress level than the true
unconfined strength. In order to achieve true unconfined behaviour,
therefore, specimens must be sized to have (L/D) ratios of at least 2.5:1.

6.6.4 Non-Parallel End Loading Effects


Preparation of core sample for testing includes machining specimens to
a symmetric size and shape. Ideally all specimens should be created to
exhibit repeatable dimensions and should be of uniform, identical size to
minimize size deviation effects. Errors occur when non-parallel end
faces are created through poor sample trimming and lathing. When
deviation occurs it is common that contact between the sample end
faces and platens can become uneven, creating areas of destress and
excessive stress. In the worst case very high point stress levels will be
built up which can result in premature failure of the rock sample. Such
failure results through formation of axial cracks, originating from the
positions of point loading, rather than by formation of true shear planar
features which are more typical of ultimate failure of rock samples in
compression.

In order to maintain proper end loading conditions, the end surfaces of


test core specimens must be cut and machined parallel

Version 1.10:Jan 2010: UNSW: GS 6-21


MINING GEOMECHANICS

6.6.5 Loading Rate and Confining Pressure Effects


The rate of loading in any test refers to either the rate of change of strain
(in strain controlled tests) or the rate of change of stress (in stress
controlled tests). The International Society for Rock Mechanics
commission recommends a loading rate or 0.5 to 10MPa/s for uniaxial
compressive tests. This corresponds to axial strain rates in the order of
10-5 to 10-4 s-1 and an attainment of peak strength in the order of 5 to 10
mins.

If we examine the strain rates used in practice:

 Explosive: 105 s-1


 Laboratory: 10-5 s-1
 Tectonic: 10-16 s-1

There are 21 orders of magnitude between quickest and slowest


application rate – for the same material.

In most rocks there is an increase in the deformation (Young‟s) modulus


and a decrease in ductility (or an increase in brittleness). The
compressive strength of rock usually increases with increase in the rate
of loading of specimens. For a range of 10-8 to 102 s-1 the compressive
strength at high rates of loading will generally be two times the
compressive strength obtained at slower rates of loading.

Archibald (2001) suggests that as the rate of specimen loading is


increased by one to two orders of magnitude, the sample failure strength
will increase by an average of approximately 10-25%.

Confining Pressure
A factor in altering the shape of the complete stress-strain curve in
compression is the effect of the confining pressure applied during the
test.

The most brittle behaviour is experienced at zero confining pressure: the


curve demonstrates less brittle behaviour (or increasing ductility) as the
confining pressure is gradually increased. At one stage the post-peak
curve is essentially a horizontal line representing continuing strain at
constant stress level (i.e. perfectly plastic). Below this line the material
strain softens; above this line strain hardening occurs. The horizontal
line is termed the brittle-ductile transition.

Confining pressures vary with rock type and can be low in some cases.
The transition represents the boundary between instability with
increasing strain (brittle behaviour) and stability with increasing strain
(ductile behaviour).

Version 1.10:Jan 2010: UNSW: GS 6-22


MINING GEOMECHANICS

In a hard rock underground mine, a creating a stope or tunnel will mean


the immediately surrounding rock mass will not be under triaxial
compression, and will therefore exhibit brittle behaviour. In deep
underground hard rock mining, we might expect more ductile behaviour
as the confining pressures increase with depth (excluding all other
factors of course). This also has implications for the different behaviour
of rock at the edge of the excavation (tunnel or stope wall, for instance)
and the same rock further removed from the edges of the excavation.

6.6.6 Environmental effects


Moisture content
The moisture content is know to influence the stress-strain curve. In this
case the mechanical response is controlled by the effective confining
pressure 3‟ = 3 – u.

u (MPa)

The example displayed here is of a series of triaxial tests on limestone


with a constant 3 = 69 MPa, but with various levels of pore pressure in
the range u = 0 to 69 MPa. There is a transition from ductile to brittle
behaviour as pore pressure is increased from 0 to 69 MPa.

The effective stress concept can apply well for materials such as
sandstone, but be inappropriate for less permeable rocks such as
granite.

Temperature and time effects


The effect of temperature of the modulus of elasticity is dependent upon
the rock type. Increasing the temperature reduces the modulus and
compressive strength, whilst increasing the ductility in the post-peak
region.

Version 1.10:Jan 2010: UNSW: GS 6-23


MINING GEOMECHANICS

6.6.7 Number of samples


After all the above influences have been accounted for in the results,
there still exists the random influence of natural flaws in sample
specimens. This normally requires a sample population of at least ten,
and more realistically twenty, specimens to produce reliable estimates of
average unconfined and triaxially-confined compressive strengths.
Typical data variation, shown below, would in the current form of a log-
normal distribution, also form input to a probabilistic model of rock
strength. This could form the basis of risk- and reliability- centred design
techniques.

30

25

20
Number

15

10

0
10

00
0

91 0

00
<5

-2

-3

-4

-5

-6

-7

-8

-9
-1
6-

>1
11

21

31

41

51

61

71

81

UCS range (MPa)

Version 1.10:Jan 2010: UNSW: GS 6-24


MINING GEOMECHANICS

References
Brady B H G, and Brown E T (1999) Rock Mechanics for Underground Mining 2nd
Ed, Chapman & Hall.
Brown, E.T., 1981 Rock Characterization Testing and Monitoring, ISRM Suggested
Methods, Pergamon, Oxford
Hoek, E. & Bray, J.W. 1997, Rock Slope Engineering, Chapman and Hall, London
Hoek, E. & Brown, E.T., 1997, Underground Excavations in Hard Rock, Chapman
and Hall, London
Hudson, J.A & Harrison, J.P. 1997 Engineering Rock Mechanics, Pergamon
Press, Oxford
Lama, R.D, & Vutukuri, V.S. 1978 Handbook on Mechanical Properties of Rocks,
Volume 2, Trans Tech Publications, Clausthal

Version 1.10:Jan 2010: UNSW: GS 6-25


Owner List
The Owners-SP 2657

Lot Unit Name and Main Contact Address Levy Address Notice Address Entitlement Balance
(-)prepaid
1 1 SAMIR BAHIG SHOUSHA 1 -1.46
2/62 MASCOT DRIVE, EASTLAKES NSW 2018 2/62 MASCOT DRIVE, EASTLAKES NSW 2018 2/62 MASCOT DRIVE, EASTLAKES NSW 2018
2 2 MR S B SHOUSHA 1 -2.45
2/62 MASCOT DRIVE, EASTLAKES NSW 2018 2/62 MASCOT DRIVE, EASTLAKES NSW 2018 2/62 MASCOT DRIVE, EASTLAKES NSW 2018
3 3 EDWARD ADHIKARI & DAISY BISWAS 1 0.00
3/62 MASCOT DRIVE, EASTLAKES NSW 2018 3/62 MASCOT DRIVE, EASTLAKES NSW 2018 3/62 MASCOT DRIVE, EASTLAKES NSW 2018
4 4 SEAN & CATHERINE O'DONOGHOE 1 0.00
4/62 MASCOT DRIVE, EASTLAKES NSW 2018 4/62 MASCOT DRIVE, EASTLAKES NSW 2018 4/62 MASCOT DRIVE, EASTLAKES NSW 2018
5 5 MR DAVID HENRY POPPLEWELL 1 0.00
5/62 MASCOT DRIVE, EASTLAKES NSW 2018 5/62 MASCOT DRIVE, EASTLAKES NSW 2018 5/62 MASCOT DRIVE, EASTLAKES NSW 2018
6 6 MR G & MRS U KASSAR 1 0.00
2 KYOGLE STREET, EASTLAKES NSW 2018 2 KYOGLE STREET, EASTLAKES NSW 2018 2 KYOGLE STREET, EASTLAKES NSW 2018
7 7 MR J & MRS M REPETTO 1 0.00
42 HIGHCLERE AVENUE, ROCKDALE NSW 42 HIGHCLERE AVENUE, ROCKDALE NSW 42 HIGHCLERE AVENUE, ROCKDALE NSW
2216 2216 2216
8 8 TRAM THI NGOG LE 1 0.00
8/62 MASCOT DRIVE, EASTLAKES NSW 2018 8/62 MASCOT DRIVE, EASTLAKES NSW 2018 8/62 MASCOT DRIVE, EASTLAKES NSW 2018
9 9 MR & MRS MAHMOUD 1 404.80
9/62 MASCOT DRIVE, EASTLAKES NSW 2018 9/62 MASCOT DRIVE, EASTLAKES NSW 2018 9/62 MASCOT DRIVE, EASTLAKES NSW 2018
10 10 YOHANES SOE & HANG LIN WIE 1 404.80
10/62 MASCOT DRIVE, EASTLAKES NSW 10/62 MASCOT DRIVE, EASTLAKES NSW 10/62 MASCOT DRIVE, EASTLAKES NSW
2018 2018 2018
11 11 MESSRS T, A & J & MRS K JAKUS 1 0.00
11/62 MASCOT DRIVE, EASTLAKES NSW 11/62 MASCOT DRIVE, EASTLAKES NSW 11/62 MASCOT DRIVE, EASTLAKES NSW
2018 2018 2018
12 12 MRS Y KASSAR 1 0.00
68 FLORENCE AVENUE, EASTLAKES NSW 68 FLORENCE AVENUE, EASTLAKES NSW 68 FLORENCE AVENUE, EASTLAKES NSW
2018 2018 2018

28/08/2009 18:49 Mark Gitman MG Strata and BMC Management Pty Ltd Page 1
Rock
MINING GEOMECHANICS

Mechanics

7
TOPIC 7:RESPONSE OF
ROCK MASS TO
UNDERGROUND
EXCAVATION

7.1 Introduction
“Knowledge of the in situ state of stress in a rock mass is essential for
the proper planning and design of mine layouts to optimise stability and
safety of mining operations. The in situ stresses are essential boundary
conditions for all design and analysis” (Stacey and Wesseloo, 2002).
Why is this? When designing surface structures, the geometry of the
structure and its operating duty define the loads imposed on the system.
For an underground rock structure, the loads imposed on the system
are generated from;

 an „initial‟ in-situ stress prior to excavation and


 an „induced‟, post-excavation state of stress

The resultant of the initial state of stress and stresses induced by


excavation define the load imposed on the system. Since induced
stresses are directly related to the initial stresses, it is clear that
specification and determination of the pre-mining state of stress is
required before any design analysis is undertaken.

The measurement and/or assessment of the in situ stress regime is


generally part of the data collection and analysis phase (refer to Figure
overleaf). However, we can also use observation and measurement
techniques as part of the monitoring process to assist in back analysis

Version 1.10:Jan 2010: UNSW: GS 7-1


MINING GEOMECHANICS

and subsequent reformulation of the mine model and modification to the


design.

Site Characterisation (Data collection and Analysis)

Mine model formulation (Engineering Analysis)

Design (Global & local)

Implementation

Rock Mass performance monitoring

Retrospective Analysis

Any undisturbed mass of rock in-situ contains non-zero stress


components due to weight of overlying materials, confinement and past
history.

Remember:
The engineering properties determined from a section of a mine may
only be considered “precise” for the location (point) of collection or
measurement. All other locations must be assumed.

The assumption of stress values at other locations is necessary


because mathematical and numerical treatments are based on
assumptions of elasticity, homogeneity and isotropy. The concept of
stress associated with a continuum (say around a partial or entire
excavation) is valid only at the scale where continuum is valid. We will
see how discontinuities and zones of low modulus can affect the
direction of the stress field and thus put the continuum concept into
question.

With a continuum assumption, the Representative Elemental Volume


(REV) is the volume of rock in which continuum assumption is valid. It is
not usually known what the scale effect outside the REV is, but you can
usually determine stress effects as either;

Version 1.10:Jan 2010: UNSW: GS 7-2


MINING GEOMECHANICS

 natural - if you are taking measurements from a greenfields site.


 induced - for all other sites adjacent to excavations

7.1.1 Learning Objectives


By the time you have completed this module you will be able to:

 Articulate the origins of in situ stress and the influencing


parameters
 Analyse the principal stress directions of a stress field
 Evaluate the stresses induced about a simple excavation
shape.
 Evaluate the influence of discontinuities on the stress field
surrounding excavations

7.2 Stress Classification


Before beginning to evaluate the response of excavations to stress, we
can start assessing the likely stress regime by gathering all the
information for the region:

1 Tectonic stress Stress state caused by tectonic plate


movement
2 Gravitational stress Stress state caused by weight of rock
above
1&2 Natural Stress The in situ stress that exists prior to
engineering
1&2 Regional Stress The stress state in a large geological
domain
1&2 Far-field stress The stress state beyond the near field

3 Local stress The stress state in a small domain

3 Near field stress The stress state in the region of an


excavation

Version 1.10:Jan 2010: UNSW: GS 7-3


MINING GEOMECHANICS

3 Induced stress The natural stress state after disruption by


excavation
4 Residual stress The locked in stress state caused by
previous tectonic activity

“Isotropic” Stress Behaviour


Stresses developed within a rock mass will not vary directionally. At any
depth below ground surface, for example, horizontal ground stresses
are assumed to be identical in magnitude in all orientations within any
horizontal plane.

“Infinite” Extent
the elemental cube model upon which ground stresses are presumed to
act lies within an infinitely large host rock matrix.

“Elastic” Material
the rock mass behaves as a purely elastic material, exhibiting elastic
stress/strain response.

The term in situ stress covers both natural and induced stresses caused
by the interaction of the excavation and the natural stress. The natural
stresses come in three forms:

 Gravitational
 Tectonic
 Active
 Remnant
 Residual.

In summary: the magnitude of the vertical stress is due to gravitational


stress and the magnitude of horizontal stress is due to tectonic stresses,
however both of these can be affected by residual stresses

In situ stresses

Natural Induced
stresses stresses
Gravitational Tectonic Residual

Active Remnant
Tectonic Tectonic
Stress Stress

Version 1.10:Jan 2010: UNSW: GS 7-4


MINING GEOMECHANICS

7.3 Horizontal and Vertical In-situ Stress


Since the ground surface is always traction-free, simple statics requires
that the vertical normal stress component (σz)(MPa) at a sub-surface
point is;

 z  r z

where γr is the rock unit weight (MN/m3), and z is the depth below
ground surface m).

Measurement of vertical stresses at different mining and civil sites


around the world confirm the relationship;

 z   v  0.027z (MPa)

is valid although there is a lot of scatter as shown in the diagram below.

For elastic rock mass behaviour in uniaxial strain („complete lateral


restraint‟) conditions, horizontal normal stress components (σh)are given
by;

  
h    v
1 

Where v is Poisson‟s ratio for the rock mass. However, this condition is
RARELY satisfied. Horizontal stress is more difficult to estimate than

Version 1.10:Jan 2010: UNSW: GS 7-5


MINING GEOMECHANICS

vertical stress. It can be estimated using the ratio of horizontal to vertical


stress, represented by k such that;

h
k
v

Hoek and Brown analysed horizontal stress data from mining and civil
engineering sites around the world and proposed:

100 1500
 0.3  k   0.5
z z

as illustrated below.

Locally, in Eastern Australia;

 Vertical stress is approximately the unit weight of rock x depth


below surface
 Horizontal stress is approximately 2 times higher than the vertical
stress in one direction and 1.5 times in the other direction

In Western Australia

 Vertical stress is approximately the unit weight of rock x depth


below surface

Version 1.10:Jan 2010: UNSW: GS 7-6


MINING GEOMECHANICS

 Horizontal stress is approximately 3 times higher than the vertical


stress in one direction and 2 times in the other direction

As will be seen later, the stress gradient and in particular the direction of
the major and minor principal stresses can significantly influence the
mining induced stresses generated around an excavation, and
ultimately effect it‟s stability.

7.3.1 Australian Stress Map


A series of projects has been undertaken to define the directions and
magnitudes of tectonic stress on the Australian continent. In the
diagram below, the stress map display the orientations of the maximum
horizontal compressive stress. The length of the stress symbols
represents the data quality, with A being the best quality. The symbol
on the line represents how the data was obtained. Methods such as
those used here are discussed in the next Topic.

Version 1.10:Jan 2010: UNSW: GS 7-7


MINING GEOMECHANICS

There is considerable variability in the in situ stress field. However it is


important to note;

1. There are three principal stresses, one oriented sub-vertically


and two oriented approximately sub-horizontally
2. The horizontal stresses are not necessarily equal and may be
different from the vertical stress
3. The ratio of average horizontal to vertical stress may be quite
high at shallow depths
4. On a continental scale, the direction of the secondary principal
stresses appears chaotic

7.4 Factors Influencing In-situ Stress

7.4.1 Topography
The effect of topography on vertical stress depends on the height of the
hill or valley in relation to its width. Mountains can increase the vertical
stress component due to the increase in overburden. Conversely
valleys will create a decrease in vertical stress but may have high
horizontal stresses. This is due to the fact that erosion causes a
decrease in depth of cover leading to a decrease in v but high k
coefficient.

7.4.2 Tectonic stresses


From elasticity theory, the horizontal stress should be:

  
h    v
1 

However, this state of stress is generally not found. The reason is that
Earth is not and has not been an inert planet. Conceptually there are:

1. Current or active tectonic stresses


2. Previous or remnant tectonic stresses and gravitational
stresses

Remnant tectonic stresses are stresses „locked‟ in (not fully relieved by


failure or deformation of the crust) from a previous tectonic regime.
Although it is difficult to identify and distinguish between 1 & 2, they can
in part account for the high stresses located in the mountainous regions

Version 1.10:Jan 2010: UNSW: GS 7-8


MINING GEOMECHANICS

of Chile and the relative high stress in the Yilgarn craton in Western
Australia.

7.4.3 Residual stresses


An example of residual stress occurs in cast iron where the exterior
cools quicker than interior, locking in stresses. A similar phenomenon
can occur with rock masses with local recrystallisation due to non-
uniform cooling. This is the biggest variable when trying to understand
in situ stress.

7.4.4 Discontinuities
Principal stresses will re-align themselves parallel and perpendicular to
an excavation surface such that there is zero stress perpendicular to the
excavation surface. In a similar fashion, the properties of both the
discontinuity and its infill material (the material contained within the
structures) affects both the strength of the structure and influences the
direction of stress. If we examine the issue in two dimensions and the
effect on the major and minor principal stress (consider the diagram
below)….

Case 1: Open discontinuity


The stresses re-align in a similar manner as they would in
any open excavation, such that the major principal is
deviated parallel with the discontinuity and the minor
principal perpendicular and has zero magnitude

Case II:Discontinuity filling has the same modulus as surrounding rock


No deviation in direction for the major or minor principal
stress.

Version 1.10:Jan 2010: UNSW: GS 7-9


MINING GEOMECHANICS

Case III: Discontinuity filling is effectively rigid


The stress field is rotated such that the major principal
stress is deviated perpendicular to the direction of the
discontinuity and the minor principal stress is deviated
parallel to the discontinuity.

7.4.5 Mining Induced in situ stress


Induced in situ stress is the stress redistribution that is created by the
interaction between the natural in situ stress field and the mining
excavation(s). There is realignment of the stress field perpendicular and
parallel to the excavations. This creates zones of stress concentration
(and increase in magnitude) and stress shadows (a decrease in
magnitude). Additionally, induced stresses can also be created via
structure loads such as large mine waste dumps or open-pit
excavations.

They can be represented by plotting the principal stress trajectories;


direction and magnitude, that reflect stress path around the excavations.
Each „cross‟ represents a major principal stress (the longer line
segment) and a minor principal stress (the shorter line segment).

Version 1.10:Jan 2010: UNSW: GS 7-10


MINING GEOMECHANICS

Consider the example of the stresses induced in the rock surrounding a


horizontal circular tunnel as illustrated in the figure below, showing a
vertical slice normal to the tunnel axis.

Before the tunnel is excavated, the in situ stresses σv , σh1 and σh2 are
uniformly distributed in the slice of rock. After removal of the rock from
within the tunnel, the stresses in the immediate vicinity of the tunnel are
changed and new stresses are induced. Three principal stresses σ1, σ2
and σ3 acting on a typical element of rock are shown above. The
convention used in rock engineering is that compressive stresses are
always positive and the three principal stresses are numbered such that
σ1 is the largest compressive stress and σ3 is the smallest compressive
stress or the largest tensile stress of the three.

The three principal stresses are mutually perpendicular but they may be
inclined to the direction of the applied in situ stress. This is evident in the
figure below which shows the directions of the stresses in the rock
surrounding a horizontal tunnel subjected to a horizontal in situ stress
σh1 equal to three times the vertical in situ stress σv (k = 3).

Version 1.10:Jan 2010: UNSW: GS 7-11


MINING GEOMECHANICS

The longer bars in this figure represent the directions of the maximum
principal stress σ1, while the shorter bars give the directions of the
minimum principal stress σ3 at each element considered. In this
particular case, σ2 is coaxial with the in situ stress σh2 ,but the other
principal stresses σ1 and σ3 are inclined to σh1 and σv in the immediate
vicinity of the tunnel.

Contours of the magnitudes of the maximum principal stress σ1 and the


minimum principal stress σ3 are given in the figure alongside. This figure
shows that the redistribution of stresses is concentrated in the rock close
to the tunnel and that, at a distance of approx three times the radius
from the centre of the hole, the disturbance to the in situ stress field is
negligible.

7.5 Stress Distributions Around Single


Excavations 1

The following section has been imported, with minor modifications, from
Sections 4 & 7 of the Deformable Materials Module (Priest, 2010) in
order to provide relevant theoretical background, and are reproduced
here with permission.

In order to understand the principles of stress analysis in three


dimensions it is first necessary to revise some of the basic
transformation principles, following which the interactive effects of
nearby excavations or discontinuities are analysed by expressing the
state of stress relative to a different set of reference axes. In this case
we describe the coordinate system as a Local Cartesian coordinate
system, and label the axes LMN. The theory has already been
developed in earlier Units.

7.5.1 Transformation of Vector Components


A vector is a quantity defined by both magnitude and orientation, such
as force or acceleration. Consider a vector V of magnitude |V|
orientated such that it makes an angle  with the positive end of the X
axis,  with the Y axis and  with the Z axis, where   and  all lie in the
range 0 to 180º. These angles are usually referred to as direction
angles.

1 These highlighted sections have already been covered in the „Deformable Materials‟ Module and are
presented here if you require them for revision and continuity between Modules

Version 1.10:Jan 2010: UNSW: GS 7-12


MINING GEOMECHANICS

X horizontal north

Vx

Vy

 Y horizontal east

V

Vz

Z vertical down

The components of V along each of the Cartesian axes are given by;

Vx = |V| cos 

Vy = |V| cos  

Vz = |V| cos 

The cosine terms above are referred to as direction cosines, as follows;

vx = cos 

vy = cos 

vz = cos 

hence

Vx = |V| vx

Vy = |V| vy

Vz = |V| vz

We can consider the same vector V in a different set of Cartesian axes,


LMN, inclined to the XYZ axes. In this inclined set of axes the
components of V are Vl Vm Vn. If we know the direction cosines of the L
axis the M axis and the N axis relative to the X,Y,Z system we can
calculate Vl Vm Vn from Vx Vy and Vz.

Version 1.10:Jan 2010: UNSW: GS 7-13


MINING GEOMECHANICS

Direction cosines relative to the


axes

Coordinate axis X Y Z

L lx ly lz

M mx my mz

N nx ny nz

We can regard V as being represented by its three separate orthogonal


component vectors Vx Vy and Vz in the XYZ Cartesian system. So Vl is
given by the sum of the components of Vx Vy and Vz in the
direction of the L axis, hence;

Vl = Vx lx + Vy ly + Vz lz

Similarly

Vm = Vx mx + Vy my + Vz mz

Vn = Vx nx + Vy ny + Vz nz

The above three equations can be written in matrix form;

 V    x y  z  Vx 
V   m 
m z  Vy 
 m  x my
 Vn   n x ny n z   Vz 

which is referred to as the transformation equation of vectors,

or [V*] = [R] [V]

The three by three matrix [R] is called the rotation matrix.

x y z 
R   m x my mz 

n ny n z 
 x

To perform the reverse rotation we can regard V as being represented


by its three separate orthogonal component vectors Vl Vm and Vn in the
LMN Cartesian system. So Vx is given by the sum of the components of
Vl Vm and Vn in the direction of the X axis, hence;

Version 1.10:Jan 2010: UNSW: GS 7-14


MINING GEOMECHANICS

Vx = Vl lx + Vm mx + Vn nx

Similarly;

Vy = Vl ly + Vm my + Vn ny

Vz = Vl lz + Vm mz + Vn nz

or

Vx   x mx n x   V 
 V    my n y  Vm 
 y  y
 Vz    z mz n z   Vn 

or [V] = [R]T [V*]

and [R]T = [R]-1

The fact that the inverse of [R] is given simply by its transpose is a
special property of the rotation matrix.

7.5.2 Transformation of Areas


The introduction to three-dimensional stress (Topic 3) was based on the
XYZ Cartesian coordinate axis system. It is, of course, permissible to
adopt any system of coordinate axes to specify the state of stress. For
example, we might wish to specify the stress state [] by reference to
the LMN Cartesian coordinate system introduced earlier

    m  n 
 *  m  mm  mn 
  n  nm  nn 

It is important to emphasise that in defining the stress state [*] the


fundamental stress state has not been changed; it is simply being
expressed relative to a different set of Cartesian coordinate axes.

In order to understand the relation between [] and [*] we must first
appreciate that in transforming stress from one set of axes to another
we are changing two things simultaneously:

(i) the components of the stresses relative to the coordinate axes,


and

(ii) the areas of the planes on which these stresses act.

Version 1.10:Jan 2010: UNSW: GS 7-15


MINING GEOMECHANICS

In the Figure below we are looking upwards towards the origin O of a


Cartesian coordinate system, with the X axis coming out of the page to
our right and the Y axis to our left (rather like looking up into the corner
of a room). A plane QRS of some general orientation and area A has
been positioned such that the normal to the plane OP is of unit length.
The direction angles of this normal are respectively ,  and  for the X,
Y and Z axes. The dimensions OQ, OR and OS of the tetrahedron
OQRS are respectively ax, ay and az.

Y
R
Q X
ay
O ax
 

Area A 
P

az S

The direction cosines of the unit vector OP are respectively

x = cos  y = cos  and z = cos 

In the right-angled triangle OPQ

OP
 cos  but OP = 1 and OQ = ax
OQ

so ax = 1/x and similarly ay = 1/y and az = 1/z

Let the areas of the right-angled triangles ORS, OQS and OQR be
respectively Ax, Ay and Az

Thus a ya z 1 a xa z 1 and
Ax   Ay  
2 2  y z 2 2 xz
a xa y 1
Az  
2 2 xy

The volume V of the tetrahedron OQRS is given by the area of the base
multiplied by one third of the vertical height, hence;

Version 1.10:Jan 2010: UNSW: GS 7-16


MINING GEOMECHANICS

Axa x 1
V 
3 6 x  y  z

But A 1
V
3

so 1
A
2  x  y z

Thus by combining the above

Ax = A x Ay = A y and Az = A z

The Figure below shows the same view of a Cartesian coordinate


system as that presented earlier, i.e. we are looking up towards the
origin O (in the „corner of a room‟). The plane QRS forms the
tetrahedron OQRS. We now consider the x components of the stresses
acting on each face of the tetrahedron, i.e. those stresses whose
second subscript is x. The stress xx acts on face ORS of area Ax, yx
acts on face OQS of area Ay, and zx acts on face ORQ of area Az.

Y
R
Q X
Az zx
A Sx
O
Ax xx
Ay yx

For static equilibrium

Ax xx + Ay yx + Az zx = A Sx

Where Sx is the x component of the stress vector acting on face QRS,


drawn reacting in the negative x direction. Similarly, in the y and z
directions allows us to define the stress vector S (sometimes referred
to as the traction vector) which has components (Sx, Sy, Sz). as;

Version 1.10:Jan 2010: UNSW: GS 7-17


MINING GEOMECHANICS

 S x   xx  xy  xz  x 
 S     
 y   yx  yy  yz   y 
 S z   zx  zy  zz   z 

Or [S] = [][]

We now consider the same stress state, but expressed relative to the
LMN Cartesian coordinate axes, reproduced below;

   m   n 
 * m  mm  mn 
  n  nm  nn 

We can envisage the cutting plane QRS with the same global
orientation as in the previous, but now cutting through the stress cube
expressed relative to the LMN axes. Proceeding in the same way as
before, we can derive the equivalent stress vector relation, as follows;

 S      m  n     
S     mm  mn   m 
 m   m
 Sn    n  nm  nn    n 

Or [S*] = [*] [*]

The vectors [S] [] [S*] and [*] like any other vectors can be
transformed from one set of Cartesian axes to another, as explained
earlier. Specifically, we define the rotation matrix [R] and adopt the
same sets of axes XYZ and LMN, so

[S*] = [R] [S]

From which;

[*] = [R] [] [R]T

This is an example of a tensor transformation equation, confirming that


mathematically the components of stress are a second order tensor.

 Rotation of vector components is a first order transformation.

Version 1.10:Jan 2010: UNSW: GS 7-18


MINING GEOMECHANICS

 Rotation of stress components involves transformation of the


equivalent stress vector components but also involves transformation
of the areas that the stresses act upon, hence it is a second order
transformation.

It is a relatively straight forward matter to substitute the full matrices in


the tensor transformation equation using;

x y z   xx  xy  xz      m  n 
R   m x my mz 

  yx  yy  yz 

 *  m  mm  mn 
n ny n z    zy  zz    n  nm  nn 
 x  zx

Multiplication of these matrices yields the following general result;

ab = axbxxx + aybyyy + azbzzz + (axby + aybx)xy + (aybz + azby)yz+


(axbz + azbx)xz

where a and b each represent one of the l, m or n direction cosines.

7.5.3 Plane problems and biaxial stress


A plane stress state exists when all the stress components with respect
to one plane in the solid are zero, e.g.

σzz = σzx = σzy=0

The stress tensor thus appears as follows:

 xx  xy 0
   yx  yy 0
 0 0 0

The example given is called a plane state of stress in the X – Y


coordinate plane. Examples of plane stress conditions in solids include
analysis and design of simple underground excavation design are based
on 3D problems idealized by 2D (biaxial) solutions. This is valid if the
excavation length to cross section dimension ratio is high. For example,
a tunnel of similar cross section can be analysed by assuming that the
stress distribution is identical in all planes perpendicular to the long axis
of the excavation.

Version 1.10:Jan 2010: UNSW: GS 7-19


MINING GEOMECHANICS

M
Y

L

X

If a set of reference axes, X,Y,Z, is established with the long axis of the
excavation parallel to the z axis, as shown above, the state of stress at
any point in the rock, for plane problems in the X Y plane, are functions
of (X, Y) only. Often, especially when considering the interactive effects
of nearby excavations or discontinuities, it is necessary express the
state of stress relative to a different set of reference axes, such as the
L,M,Z axes shown above. The direction cosines of new (L, M), relative
to the old (X, Y) axes are thus;

Direction cosines relative to the axes


Coordinate axis X Y
L lx=cosα ly=sinα
M mx=-sinα my=cosα

Introducing these values into the general form of the transformation


equations (biaxial stress when the Z-axis is a principal axis), the obverse
shear stress components vanish which gives;

 ll   xx cos2    yy sin2   2 xy sin cos

 mm   xx sin2    yy cos2   2 xy sin cos

 lm   xy cos2   sin2     xx   yy sin cos

Using the identities;

sin2  
1
1  cos2 , cos 2 
1
1  cos2 
2 2

Version 1.10:Jan 2010: UNSW: GS 7-20


MINING GEOMECHANICS

 ll 
 xx   yy 

 xx   yy cos 2
  xy sin 2
2 2

 mm 
 xx   yy 

 xx   yy cos 2
  xy sin 2
2 2

 lm   xy cos 2 
 xx   yy sin 2
2

Note that;

 xx   yy   ll   mm

2 xy
 lm  0 when tan 2 
 xx   yy

These results have special significance, as discussed below.

For any general state of stress it is possible to implement a stress


transformation such that each of the resulting shear stresses is zero. In
other words, it is possible to rotate the stress cube to find a single
unique orientation such that the shear stresses on each face are zero.

The direction given by the value of α in the above relationship when


σlm = 0 is known as a principal direction, because in this direction the
only non-zero tensor components are normal stresses. The faces then
become principal planes and the normal stress on each face is called
a principal stress. The line of action of each principal stress is called a
principal axis. (For example, when we stand still on slippery ice, which
cannot provide any significant shear reaction, we must ensure that the
interface between our shoes and the ice is a principal plane, otherwise
we will slip over. We achieve this goal by standing vertically.)

When there are three normal stresses there will be three principal
stresses, as follows:

 1 is the major principal stress, which is the largest (or least


negative) stress
 2 is the intermediate principal stress
 3 is the minor principal stress, which is the smallest (or most
negative) stress

In the case of plane geometry or biaxial stress, there are two normal
stresses so there will be two principal stresses, as follows

Version 1.10:Jan 2010: UNSW: GS 7-21


MINING GEOMECHANICS

 1 is the major principal stress, which is the largest (or least


negative) stress
 2 is the minor principal stress, which is the smallest (or most
negative) stress

Hence, in two dimensions, four bits of information are necessary,


namely the values of the two principal stresses σ1 and σ2 together with
their angles α1 and α2. The angles must be 90 degrees apart for the
standard Cartesian coordinate system.

The analysis of stresses adjacent to an excavation will usually produce


a general stress tensor (where the shear stresses are non-zero)
expressed relative to a local set of axes. Strength models for the rock
are usually expressed in terms of the principal stresses, so it is important
to be able to calculate the values and orientations of the principal
stresses associated with some general stress tensor in order to assess
rock stability.

It can be shown that for the case of plane geometry or biaxial stresses;

 1, 2 
 xx   yy 
    yy   4 xy
2 2

xx
2

Note that

 xx   yy   ll   mm   1   2

The orientations of the respective principal stress axes are obtained by


establishing the direction of the outward normal to a plane which is free
of shear stress, which gives;

1  2 xy 
1, 2  tan 1  
2   
 xx yy 

By differentiating σlm with respect to α, and setting the result equal to


zero, we can find the angle at which σlm reaches its maximum value.
This angle is shown to be 45 from the principal stress directions. This
is a special shear stress and is denoted by τmax, and is given by:

 max 
 1   2     xx   yy  
2   2 xy 

2  2  

Version 1.10:Jan 2010: UNSW: GS 7-22


MINING GEOMECHANICS

7.5.4 Cartesian and Polar Co-ordinates


In two-dimensional analysis of axially symmetrical structures it is
generally convenient to represent the state of stress and strain in terms
of polar coordinates r .

 The angle  (0 to 2) gives the anti-clockwise angle of rotation from a


reference axis.
 The distance r gives the radial distance from the centre of the axially
symmetrical structure.

In two dimensions the normal stress components are rr (radial stress)
and  (tangential or circumferential stress) with the associated
shear stress r. Two-dimensional polar coordinates can be extended to
three-dimensional cylindrical coordinates by defining the Z axis to
correspond to the axis of the axially symmetrical structure. In this case
the additional normal stress component is zz (axial stress) and the
associated shear stresses are rz and z. Two-dimensional and three-
dimensional strains can be specified in the same way.

It is important to remember that the actual orientation of the radial and


tangential stress or strain components will vary with location around the
axially-symmetrical structure. Polar and cylindrical coordinate systems
are adopted because the states of stress and strain around axially
symmetrical structures generally vary directly as functions of the
parameters r and . The parameters r and  therefore not only serve to
specify the orientation and location of the stress/strain components, but
also appear in functions that define the values of the stress/strain
components themselves at the specified location.

The relationship is shown below:

Version 1.10:Jan 2010: UNSW: GS 7-23


MINING GEOMECHANICS

7.6 Analysis of Stresses Around Underground


Excavations
The following section has been imported, with minor modifications, from
Sections 15 &16 of the Deformable Materials Module (Priest, 2010) in
order to provide relevant theoretical background, and are reproduced
here with permission.

The design of an underground excavation generally follows the following


broad strategy, as illustrated below:

1. An initial design is developed that satisfies the need to gain


access to, excavate or ventilate the ore body
2. The stresses in the rock immediately adjacent to the
excavation wall are estimated by means of analytical or
numerical methods
3. If the predicted stresses are found to be greater than the rock
mass strength then further calculations are required to identify
the extent of the zone of failed rock. If the failure zone is
extensive then it will be necessary to modify the excavation
design or to design a suitable rock support system.
4. If the predicted stresses are found to be less than the rock
mass strength, then it is necessary to assess the likely impact
of mechanical discontinuities in the rock mass. In particular,
the likelihood of slip along a major discontinuity, or the failure
of discrete rock blocks, need to be assessed.

Version 1.10:Jan 2010: UNSW: GS 7-24


MINING GEOMECHANICS

Estimation of the stresses adjacent to an underground excavation is


usually based on numerical modelling techniques using boundary
element or finite element packages. These numerical modelling
methods are discussed in more detail in Geotechnical Engineering. A
number of valuable insights into such matters as localised rock yield,
stress concentration and zones of stress influence can, however, be
obtained from analytical, closed-form solutions such as those outlined
here. In addition, these closed form solutions can be used to check and
validate the results produced by numerical models.

7.6.1 Tunnel or Shaft of Circular Cross Section


Published in 1898 by Kirsch for the simplest cross-sectional shape, the
circular hole. Consider a tunnel the cross section of radius a, driven in
the direction z. Take the rock mass in which the tunnel is driven to be of
infinite extent, and suppose that rr = θ = 0 at r = a and rr = Kσyy for θ=
0 and r = 0 at r = ∞.

The radial, tangential and shear stresses at an anti-clockwise angle θ


measured from the X axis, and a radial distance r from the centre of the
hole are given by:

ur u
rr
yy
r


r

a

xx = K yy X

 yy   a2   3a 4 4a 2  
 rr  1  K 
 1  
2 
 1  K 1  4  2  cos 2 
2   r   r r  

Version 1.10:Jan 2010: UNSW: GS 7-25


MINING GEOMECHANICS

 yy   a2   3a 4  
   1  K 1  2   1  K 1  4  cos 2 
2   r   r  

 yy   2a 2 3a 4  
 r   1  K 1  2  4   sin 2
2   r r  

On the boundary (r = a, rθ = 0,rr = 0 );

   yy 1  K   21  K cos 2 

This observation is important because rock failure will always be initiated


at the point where the stresses are at their most deviatoric, i.e. adjacent
to the wall of the excavation.

Tensile or compressive failures may occur on the boundary


according to the following conditions;

σθθ ≥ UCS – compressive failure

σθθ ≤ UTS – tensile failure

i.e;

 yy 1  K   21  K cos 2   UCS

 yy 1  K   21  K cos 2   UTS

These equations can be re-arranged to give the angles at the


boundary of a tunnel where stable, compressive and tensile failure
may be expected given a specific K ratio.

For the case where σxx=σyy, i.e. K = 1 (hydrostatic stress conditions), σrθ
= 0 and;

 a2 
 rr  yy 1  
 r 2 

 a2 
   yy 1  
 r 2 

Version 1.10:Jan 2010: UNSW: GS 7-26


MINING GEOMECHANICS

For this two-dimensional case, an internal radial pressure pi inside a


circular hole of radius a, will theoretically generate at a radial distance r
from the centre of the hole radial and tangential stresses (σrθi = 0) given
by the following expressions:

pi a 2
 rr i 
r2

pi a 2
  i  
r2

In the current context, this internal pressure pi can be regarded as a


support pressure generated by rock bolts, shotcrete or some other
support system. Support pressures in underground excavations would
normally be less than about 0.2 MPa.

If we let the radial and tangential shear stresses predicted by the Kirsch
equations be respectively rrk and r we can add these stresses to
those generated by internal support pressure pi by simple superposition
as follows:

 rr   rrk   rri

    k   i

 r   rk   ri

The stability of the rock mass subjected to the stresses in predicted in


these equations will depend on its strength, as discussed in Topic 5.

The stress at the tunnel wall (r = a) (tangential stresses), or the


boundary conditions, varies as shown below;

Version 1.10:Jan 2010: UNSW: GS 7-27


MINING GEOMECHANICS

yy

σθθB

σθθA
r
a

xx = Kyy X

For K<1.0

At point A;  A   A   yy 3  K 

At point B;  B   B   yy 3K  1

For K = 0;

At point A;  A   A  3 yy

At point B;  B   B   yy

Note that tensile boundary (tangential) stresses can thus exist in a


compressive stress field.

For K = 1;

At point A and B;  A B   AB  2 yy

Implications of Calculated Stress on Yield Strength Models


Assuming there is no internal pressure, the radial stress is zero at the
excavation boundary, so the rock is in a state of uniaxial compression
and its yield strength is given by the uniaxial compressive strength of the
rock.

Version 1.10:Jan 2010: UNSW: GS 7-28


MINING GEOMECHANICS

The Coulomb (strength) criterion shows that the yield strength of rock is
substantially increased by the effect of confining pressure. For example,
a rock with a cohesion of 1.4 MPa and an angle of friction of 35 has a
theoretical uniaxial compressive strength of only 5.379 MPa. A
confining pressure of 0.12 MPa, generated by a support pressure on the
boundary of the excavation, can be represented by setting 3 to 0.12
MPa in the Coulomb criterion, giving a yield strength of 5.822 MPa,
which is an increase of some 0.443 MPa or nearly four times the
confining stress.

This simple analysis demonstrates that the action of rock support is, in
reality one of generating a small confining pressure which dramatically
increases the strength of the rock, allowing it better to withstand the
stress concentrations adjacent to the excavation boundary. If we accept
this model then it is more appropriate to call rock bolts and shotcrete
„rock reinforcement‟ rather than „rock support‟ because they strengthen
the rock rather than supporting it like columns in a building.

When K = 0 the stress concentration factors are 3 and –1 respectively,


implying that tensile stresses will develop at the  = 90 location. A
stress field with K = 0 is unlikely to occur deep within a rock mass, but in
regions adjacent to a large free surface or underground excavation, the
stress in one direction may become relieved almost down to zero and
thereby create a K = 0 stress field. In such a stress field, the creation of
tensile stresses at the  = 90 location in a circular excavation is a
serious matter. Rock masses are not good at sustaining tensile
stresses, so yield and cracking is likely to occur under these conditions.
If it is the horizontal stress that has been relieved to zero, the  = 90
location will be in the roof of the excavation, so rock yield will be
compounded by the effects of gravity; indeed engineers may, perhaps
erroneously, ascribe the cause of the rock failure to the gravitational
dislocation of rock blocks in the tunnel roof.

The previous analysis is based on a two-dimensional analysis. A


pseudo-three-dimensional analysis can be evaluated by considering the
stress zz parallel to the axis of the circular opening. The minor principal
stress 3 is taken to be the smallest, or most negative of {rr  zz}.
The major principal stress 1 is taken to be the largest, or most positive
of {rr  zz}. In assigning these values we are ignoring the fact that r
is not necessarily zero, and that rr and  are not, therefore, strictly
principal stresses. The value of 3 can be input into the Coulomb or
Hoek-Brown strength models to determine the value of 1 at failure,
here termed 1f. This value of 1f is then compared to the computed
value of the major principal stress 1 at the specified point within the
rock mass.

 If 1f > 1 then the rock is predicted to be stable

Version 1.10:Jan 2010: UNSW: GS 7-29


MINING GEOMECHANICS

 If 1f < 1 then the rock is predicted to yield

Although localised zones of yielding can be tolerated, substantial zones


of rock mass yield would necessitate a re-design of the excavation
geometry and/or the support system.

There are a number of yield criteria that take each of the three principal
stresses into account when calculating rock strength. Consideration of
these three-dimensional yield criteria is beyond the scope of this course.

7.6.2 Zones of Stress Influence


The stress perturbation caused by excavation decreases with distance
from the wall of the excavation. With K=1, the radial and tangential
stresses, expressed as a ratio of the remote in situ stresses are as
follows;

 rr  a2 
 1  2 

 yy  r 

   a2 
 1  2 
 yy  r 

The radial stress perturbation ratio is –(a2/r2) and the tangential stress
perturbation ratio is +(a2/r2). For example, when a/r is 0.5 then r = 2a, so
we are looking at a point one excavation radius beyond the excavation
wall. At this point the stress perturbation will be (0.5)2 or 25%. When a/r
is 0.2 we are looking at a point four excavation radii beyond the
excavation wall; at this point the stress perturbation is only 4%. This
point of 4% stress perturbation, two excavation diameters from the
excavation wall, is generally regarded as the limit of influence of the
stress concentration induced by a circular, or approximately circular,
excavation.

Clearly, the actual size of the zone of stress influence will depend on the
size of the excavation. The zone of stress influence of a 30 m „diameter‟
stope will extend to some 60 m beyond the wall of the stope. A 5 m
diameter development drive located 20 m from the stope will therefore
be driven through rock subjected to stress concentration effects of the
stope, rather than the „pristine‟ in situ stress field. The stope will,
however, be outside the zone of stress influence of the development
drive. This development drive will create its own additional stress
concentration within this distorted stress field, which could lead to
substantial rock yield. This situation will be exacerbated if the stope is
extended and backfilled. The zone of stress influence may well expand

Version 1.10:Jan 2010: UNSW: GS 7-30


MINING GEOMECHANICS

and retreat, exposing the development drive to the damaging effects of


cyclic loading and blast vibrations.

A pair of 30 m diameter excavations separated by 80 m will be outside


each other‟s zones of stress influence. There will, however, be a zone
of rock some 40 m thick that lies within the zones of stress influence of
both excavations. To calculate the stress state in this zone of overlap,
we need to consider the stress increments produced by each
excavation. The following example illustrates this procedure.

yy = 30 MPa

30 m 80 m 30 m

xx = 40 MPa

zz = 35 MPa
Stope A Stope B

The important point to note is that, since we are considering the stress
state along a line drawn between the centres of the two „circular‟ stopes,
their radial stresses will be horizontal and parallel to the X Cartesian
coordinate axis, and their tangential stresses will be vertical and parallel
to the Y Cartesian coordinate axis. This property would not apply along
any other direction. The coincidence of the radial and horizontal
stresses, and the tangential and vertical stresses allows us to add and
subtract stress increments directly. This process, which is known as
superposition, can only be done when the component stresses are
parallel. In order to consider stress increments along some other axis
we would need to undertake a two-dimensional stress transformation to
calculate the xx and yy stress components from {rr  r} for each
stope.

At each point, the stress increment in the x direction xx produced by


the excavation of stope A is given by rr  xx; similarly the stress
increment in the y direction yy is given by   yy. Similar
calculations for each point give the stress increments produced by the
excavation of Stope B, xx and yy. The final stress state at each
point is given by the algebraic sum;

xx + xx + xx and yy + yy + yy.

Version 1.10:Jan 2010: UNSW: GS 7-31


MINING GEOMECHANICS

7.6.3 Effects of Planes of Weakness on Stress Distribution


In excavation design problems where major discontinuities are located
in the vicinity of the tunnel, in some cases, an elastic analysis is a
suitable basis for design in a discontinuous rock mass, and in others,
provides a reasonable indication of likely effect of a discontinuity.

In this analysis, a low discontinuity shear strength (and zero tensile


strength) defined by the Mohr Coulomb model is assumed compared
with that of the intact rock mass.

The example here is based on a circular cross section, but, the


principles apply to an opening of other shapes also (if the stress
distribution around the opening is known).

rr
 x yy
X

r 
d

r

a

xx = yy X

σn
τ
Y

A circular opening in a hydrostatic stress field (K=1) is to be excavated


close to, but not intersecting, a plane of weakness. From the geometry
given above and the equations for the stress distribution around a
circular hole and the transformation equations, the normal and shear
stresses on the plane are given by;

n 
 rr       rr     cos 2  a2 
  yy 1  2 cos 2 
2 2  r 

   r
    sin 2
cos 2  rr 
 yy a 2 sin 2
2 r2

Version 1.10: Jan 2010: WASM:RJT 7-32


MINING GEOMECHANICS

The value of the ratio shear/normal stresses determined above is plotted


for various points along the plane of weakness from the centerline of the
opening.

The peak value of the shear stress/normal stress ratio corresponds to


the mobilized angle of friction. If the angle of friction for the plane of
weakness exceeds this value, no slip is predicted on the plane, and the
elastic stress distribution can be maintained. For a plane of weakness
with an angle of friction of less than the mobilized value, the extent of the
predicted zone of slip is shown. Since the problem is symmetric about
the opening centerline, a zone of slip is also predicted on the other side.
For both zones, the sense of slip produces inward displacement of rock
on the underside of the plane of weakness. This would be expressed as
increased boundary stresses in the segment between the fault and the
excavation. The effect of the fault is to deflect and concentrate the stress
trajectories in the region between the excavation and the fault.

7.6.4 Tunnel or Shaft of Ellipsoid Cross Section


A comprehensive, complex solution is available for all points around
an elliptical excavation, but in this treatment, only boundary stressed
(surface tangential) stresses are considered.

B
H = 2b A Kp

W = 2a

b2 a2 a
A  B  q
a b b

From theoretical development,

Version 1.10:Jan 2010: UNSW: GS 7-33


MINING GEOMECHANICS

 2w 
At point A  A   A  p1  K  2q   p1  K 
  A 

 2K   2H 
 B   B  p  K  1   p K  1  K 
q 
At point B 
   B 

References
Amadei, B and Stephanssom, O. 1997 Rock Stress and its Measurement,
Chapman and Hall, Cambridge
Brady B H G, and Brown E T (1999) Rock Mechanics for Underground Mining 2nd
Ed, Chapman & Hall.
Goodman R E (1989) Introduction to Rock Mechanics, Ed, John Wiley & Sons.
Hudson J A & Harrison J P (1997) Engineering Rock Mechanics, An Introduction to
the principles, Pergamon.
Priest S. D. (2010). Deformable Materials Module lecture notes, Mining
Geomechanics. The University of Adelaide.

Version 1.10:Jan 2010: UNSW: GS 7-34


Rock
MINING GEOMECHANICS

Mechanics

8
TOPIC 8: METHODS OF
IN-SITU STRESS
MEASUREMENT

8.1 Introduction
In the previous Topic, the importance of determining the pre-mining or
concurrent with mining state of stress was emphasized, leading as it
does into excavation design, analysis of excavation stability and
ultimately support design.

Methods developed to achieve this generally require access to the area


of interest, either by personnel, or more cost-effectively, by use of a
borehole. Thus, most of the methods of in situ stress measurement
involve the observation of a change in deformation or stress resulting
from a change in the geometry of an opening in the rock, and the
subsequent calculation of the field stresses from those measured
changes.

The determination of in situ stress can be categorised via:

 Observational techniques
 Measurement techniques – of which the latter has three types:
 Destressing Techniques
 Destressing – Restressing Techniques
 Overstressing Techniques

The difference between the techniques is shown (conceptually) below.

Version 1.10: Jan 2010:WASM:RJT 8-1


MINING GEOMECHANICS

In-situ
stress

De- De-
stressing stressing
methods re-
stressing
methods

Destressing techniques involve relieving the stresses by wholly or


partially isolating the sample volume from the in situ stress field.

Destressing-restressing techniques involve wholly or partially relieving


the stresses and then returning the sample volume to its original stress
state.

Overstressing techniques involve stressing the rock material until the in


situ material strength is exceeded.

The most common technique is based on determination of strains in the


wall of a borehole, or other deformations of the borehole, induced by
overcoring (de-stressing) that part of the hole containing the
measurement device.

The second type of procedure, flatjack measurements and hydraulic


fracturing, determines a circumferential normal stress component at
particular locations in the wall of a borehole. At each location, the
normal stress component is obtained by the pressure, exerted in a slot
or fissure, which is in balance with the local normal stress component
acting perpendicular to the measurement slot. The circumferential
stress at each measurement location may be related directly to the state
of stress at the measurement site, preceding boring of the access hole.
This is a typical de-stress re-stress procedure.

The third method is based on the analysis and interpretation of patterns


of fracture and rupture around deep boreholes such as oil and gas wells.

Version 1.10:Jan 2010: UNSW: GS 8-2


MINING GEOMECHANICS

8.1.1 Learning Objectives


By the time you have completed this module you will be able to:

 Contrast recognized techniques for measuring in situ stress


 Describe a new techniques for measuring in situ stress
 Determine the magnitude and directions of the principal stress
directions from data derived from various stress measurement
techniques.

8.2 Methods of Stress Measurement

8.2.1 Observational Techniques


Observations made of openings or boreholes can provide valuable
indications of the magnitudes and orientations of in situ stress. These
techniques fall into the category of “overstressing”.

Borehole breakout
Borehole breakout refers to the stress induced failure that occurs in the
walls of a borehole. The locations of the breakout (on diagonally
opposite sides of the borehole) are usually aligned with the intermediate
principal stress acting in the plane normal to the borehole axis. They
can provide a reliable indication of the orientation of the in situ stress
fields.

 



Version 1.10:Jan 2010: UNSW: GS 8-3


MINING GEOMECHANICS

Core discing
Core discing appears to be closely associated with the formation of
borehole breakouts. In brittle rock, discing and breakouts occur over the
corresponding lengths of core and borehole. The thinner the discs the
higher the stress level.

A measure of the inclination of the principal stress to the borehole can


be gauged from the relative asymmetry of the disc. For unequal stress
normal to the core axis, the core circumference will peak and trough as
shown below. The direction of the line drawn between the peaks of the
disc surfaces facing in the original drilling direction indicates the
orientation of the intermediate principal stress.

8.2.2 Destressing Techniques


CSIRO Hollow Inclusion (HI) Cell
The overcoring technique is a popular method of rock stress
measurement and the CSIRO Hollow Inclusion Cell (HI Cell) is used for
this technique. It involves measuring changes, deduced from readings

Version 1.10:Jan 2010: UNSW: GS 8-4


MINING GEOMECHANICS

from strain gauge rosettes (A, C B), induced by creating a cylindrical


surface around an installed Hicell.

Stress relief in the vicinity of the strain cell induces strains in the gauges
of the strain rosettes, equal in magnitude but opposite in sign to those
originally existing in the borehole wall. The state of strain in the wall of
the borehole prior to stress relief can then be established.

Version 1.10:Jan 2010: UNSW: GS 8-5


MINING GEOMECHANICS

Installation Overcoring (destressing)

The steps used to determine stress with this instrument are:

1. A large diameter borehole is drilled beyond the zone of


influence of the excavation
2. A smaller diameter pilot hole is extended from the end of the
larger hole
3. The Hi Cell is installed into the pilot hole within an intact interval
of rock
4. An initial set of strain readings are taken
5. The larger borehole is then extended over the pilot hole,
effectively destressing it from in the in situ stress field
6. Strain readings are taken during the overcore drilling and after
the overcoring is complete
7. The state of strain is calculated from the difference in the strain
readings before and after overcoring
8. The large overcore containing the inclusion is recovered and
subject to increasing biaxial pressures in a pressure cell.
During this test strain readings are recorded from the inclusion
9. The elastic properties of the core are calculated using the strain
changes induced by the test pressures during biaxial testing
10. An estimate of the state of the in situ stress is made using the
relieved state of strain and the assumed constitutive
relationship based on the elastic properties of the rock

The main advantages of the technique are full 6-component stress


tensor measurements and 3 to 4 rosettes allow for redundant values
and statistical analysis of results. However, the instrument is generally
unrecoverable and cannot be relocated to another test location.

Version 1.10:Jan 2010: UNSW: GS 8-6


MINING GEOMECHANICS

8.2.3 Destressing - Restressing Techniques


Flatjack
Stress measurement using strain gauge devices asses the state of
stress over small volumes. Larger volumes of rock can be examined if a
larger diameter opening is used as the measurement site. For openings
allowing human access, it may be more convenient to measure directly
the state of stress in the excavation wall, rather than the state of strain.
This eliminates the need to determine or estimate a deformation
modulus for the rock mass.

The basic principle of the flatjack is shown below. It requires a large


excavation (tunnel or stope). Two pins are drilled and fixed into the
excavation boundary and the distance (do) is measured between them.
A slot is cut into the rock between the pins. If the normal stress is
compressive the pins will move together as the slot is cut. A flatjack,
(comprised of two metal sheets welded together around the edge with a
feeder tube attached), is grouted into the slot. The flatjack is
pressurized with water or oil and the pins move apart. It is assumed that
the pressure (cancellation pressure) required to move the pins back to
the original position is the same as the pre-existing normal stress.

Three prerequisites must be satisfied for a successful in situ stress


determination using flatjacks. These are:

 a relatively undisturbed surface of the opening constituting the


test site;
 an opening geometry for which closed-form solutions exist,
relating the far-field stresses and the boundary stresses
 a rock mass which behaves elastically, in that displacements are
recoverable when the stress increments inducing them are
reversed.

The first and third requirements virtually eliminate the use of a test site
excavated by conventional drilling and blasting, which will cause
extensive disturbance of the elastic stress distribution in the rock and
may give rise to non-elastic displacements in the rock during the

Version 1.10:Jan 2010: UNSW: GS 8-7


MINING GEOMECHANICS

measurement process. The second requirement restricts suitable


opening geometry to simple shapes. An opening with circular cross
section is by far the most convenient. A test only produces one normal
stress, thus a minimum of 6 tests in independent directions are required
to obtain full stress tensor. 9 tests are usually performed; 3 each in
backs, sidewall and face.

8.2.4 Overstressing Techniques


Hydraulic Fracture
The method involves the pressurization of a length of borehole and the
measurement of the pressure required to fracture the rock or re-open
existing fractures.

Conventional hydraulic fracturing involves the pressurizing of a short


length of borehole, isolated using hydraulic packers on either side of it,
until the hydraulic pressure causes the rock to fracture. The
characteristics of the pressure induced breakdown and the subsequent
reopening of the fracture under repressurisation are monitored carefully.
The orientation of the induced fracture is measured using a borehole
television camera or a special impression packer to obtain a physical
record of the surface of the borehole. From all these data the
orientations of the secondary principal stresses normal to the axis of the
borehole can be interpreted. Vertical boreholes are usually used and it
is assumed that the in situ principal stresses are vertical and horizontal.
Hydraulic fracturing is a well established method of in situ stress
measurement, particularly in the oil well industry.

To calculate the minimum horizontal principal stress, the borehole is


vertical and parallel with the principal stress component in the vertical
direction. Thus the calculation only deals with stresses in the horizontal
plane.

 h  ps
Where ps = pressure to equilibrate fracture-normal stress, h in a
direction normal to fracture (shut-in pressure).

To calculate the maximum horizontal principal stress (H),

 H   t  3 h  pb
Where t = tensile strength of rock and pb = breakdown pressure (crack
re-opening).

Version 1.10:Jan 2010: UNSW: GS 8-8


MINING GEOMECHANICS

Hydraulic Tests on Pre-existing Fractures (HTPF)


The HTPF method is a variation on conventional hydraulic fracturing. It
consists of reopening an existing fracture that had previously been
identified in the borehole and isolated with packers. The orientation of
the fractures also needs to be measured. The pressure that just
balances the normal stress acting across the fracture is measured. By
identifying at least six non-parallel pre-existing fractures and carrying out

Version 1.10:Jan 2010: UNSW: GS 8-9


MINING GEOMECHANICS

hydraulic tests on such fractures, six components of in situ stress can be


determined, and hence the three dimensional in situ state of stress can
be interpreted.

Its main advantages are that the borehole does not have to be assumed
to be drilled in a principal stress direction and it requires no assumptions
regarding the in situ stress conditions or rock mass deformation and
strength properties. Neither is it necessary to determine the tensile
strength of the rock. However, identifying the 6 non-parallel existing
fractures is difficult and the resulting calculations are complex.

8.2.5 Kaiser Effect


Kaiser (1953) observed that when the stress on a polycrystallised metal
was relaxed and then reapplied, there was a significant increase in the
rate of acoustic emission when the previous maximum stress level was
exceeded. This phenomenon has become known as the Kaiser effect
and Goodman (1963) observed a similar effect in rocks. Research was
undertaken in Japan during the 1970’s to make use of the phenomenon
to estimate in situ stresses and a considerable amount of further
development of the method has taken place including recent work by
Villaescusa et al (2003) at the Western Australian School of Mines,

The Kaiser effect method involves the drilling of small secondary,


orientated cores from the original core removed from the stressed
environment.

The original core must be orientated so the directions of secondary


coring are known in relation to this original core orientation. The
secondary cores are prepared with the required end flatness and

Version 1.10:Jan 2010: UNSW: GS 8-10


MINING GEOMECHANICS

parallelism, and then subjected to uniaxial compressive stress whilst the


acoustic emissions from the rock are monitored using sensors attached
to the core.

On a plot of the applied stress vs. the acoustic emissions, the Kaiser
effect change point is at the position on the curve where the slope of the
plot noticeably increases. As the Kaiser effect changes in acoustic
emission rate, the stress corresponds with the previous maximum stress
to which the rock had been subjected.

If a sufficient number of secondary cores are tested, the full three


dimensional in situ state of stress may be determined. The advantages
of the technique are;

Version 1.10:Jan 2010: UNSW: GS 8-11


MINING GEOMECHANICS

1. It gives a direct measure of stress. It is not dependent on the


measurement of strain and the subsequent calculation of stress
from strain, which requires the assumption of a relationship
between stress and strain for the rock as well as measurement
of the deformation properties of the rock. All of these factors
can introduce errors.
2. The full three dimensional in situ state of stress may be
determined.
3. Use can be made of original core obtained for other purposes,
such as exploration, making the method cost effective.
4. Core obtained remotely can be used, and therefore the method
is applicable to Greenfield sites, before any excavations have
been made, as well as to operating mines.
5. Since small secondary cores are used for the tests, many tests
can be carried out using a limited length of original borehole
core. Again this makes the method cost effective, with a large
number of results being able to be obtained at relatively low
cost. The more the number of cores tested, the greater the
confidence in the results obtained .
6. The necessary accurate and sensitive sample preparation and
testing activities are carried out in the laboratory. This therefore
obviates field based testing errors that are common in the often
harsh mining environment. Interference with production mining
operations is also reduced or eliminated.

References
Amadei, B and Stephanssom, O. (1997) Rock Stress and its Measurement,
Chapman and Hall, Cambridge
Brady B H G, and Brown E T (1999) Rock Mechanics for Underground Mining 2nd
Ed, Chapman & Hall.
Goodman R E (1989) Introduction to Rock Mechanics, Ed, John Wiley & Sons.
Villaescusa, E., Windsor, C.R., Li, J., Baird, G., Seto, M, (2003). Stress
Measurement from Cored Rock, Project M329, Minerals and Energy Research
Institute of Western Australia,
Villaescusa, E., Seto, M, Baird, G., (2002). Stress Measurement from oriented
cored, Int Jnl Rock Mechanics and Mining Sciences, 39, 603-615
Hudson J A, Cornet, F.H & Christiansson, R.(2003) ISRM Suggested Methods for
rock stress determination – Part 1: Strategy for rock stress estimation, Int Jnl Rock
Mechanics and Mining Sciences, 40, pp 991-998
Sjoberg, J., Christiansson, R & Hudson J A, & Cornet, F.H., (2003) ISRM
Suggested Methods for rock stress determination – Part 2: Overcoring Methods,
Int Jnl Rock Mechanics and Mining Sciences, 40, pp 999-1010
Haimson, B.C., & Cornet, F.H. (2003). ISRM Suggested Methods for rock stress
determination – Part 3: Hydraulic fracturing and/or hydraulic testing of pre-existing
fractures, Int Jnl Rock Mechanics and Mining Sciences, 40, pp 1011-1020

Version 1.10:Jan 2010: UNSW: GS 8-12


Rock
MINING GEOMECHANICS

Mechanics

9
TOPIC 9: TIME
DEPENDANT AND
DYNAMIC BEHAVIOUR
OF ROCK

9.1 Introduction
Previously we have examined elastic and inelastic behaviour and
mentioned that time is one of the parameters that influence
performance. Time-dependent (TD) behaviour of a rock occurs
whenever any one of the state variables, like stress, strain, deformation,
heat, flow, etc., changes with time. Earthquake loading, blasting,
transient water and heat flow, consolidation (dissipation of pore water
pressure through a loaded rock) are examples of time-dependent
problems.

As an example of TD, deformation or strain caused by a constant stress,


called creep, is like secondary consolidation (creep at a fixed effective
stress) in consolidation tests of soils or soft rocks.

Creep can initiate from the intact rock or discontinuities (fissures/ flaws/
joints). Rock creep is temperature sensitive which may imply that rock
creep is basically a thermally activated process. It means a change in
temperature causes thermal strains leading to more deformation,
fractures or fracture-enlargements and finally rock failure.

Before we continue though, let us review the general rock engineering


process and determine where this topic fits.

Version 1.10: Janr 2010:WASM:RJT 9-1


MINING GEOMECHANICS

Site Characterisation (Data collection and Analysis)

Mine model formulation (Engineering Analysis)

Design (Global & local)

Implementation

Rock Mass performance monitoring

Retrospective Analysis

Time dependent behaviour should theoretically be part of step 2,


however it is very difficult to predict the behaviour in practice and is
therefore generally tackled as a part of monitoring and retrospective
analysis:

 Stage 5: Monitoring movements. This will generally be the first


area to establish TD behaviour. Retrospective analysis will then
lead to…..
 Stage 2: The relevant formulations need to be reviewed. This
may be a quick or more in-depth, partly depended on the rate of
the TD behaviour.
 Stage 3: If the TD behaviour is relatively rapid, an immediate
response may be to redesign reinforcement and support.

9.1.1 Learning Objectives


By the time you have completed this topic you will be able to:

1. Differentiate between time dependent empirical laws and


rheological models & laws based on physical processes
2. Describe common time-dependent rheological models and their
elements
3. Calculate time-dependent behaviour of intact rock from
laboratory data
4. Calculate time-dependent behaviour of rock defects from
laboratory data

Version 1.10:Jan 2010: UNSW: GS 9-2


MINING GEOMECHANICS

9.2 Basic Concepts


Time is not taken into account in the theory of elasticity. When loading
or unloading a rock sample, it is assumed stresses and strains develop
instantaneously. However, the complete stress-strain curve character is
influenced by strain rate. Since rock continues to deform after a stress
change (for instance, due to excavation), then whilst the theory of
elasticity helps in understanding the mechanics of rock masses, a theory
for TD is also required.

In this Topic, content has been developed from Malan (1988, 2006)and
Dusseault & Fordham (1993).

Excavation in a rock mass changes the stability of the rock mass and
the subsequent readjustment of the rock mass towards a new
equilibrium does not occur instantaneously but as a gradual process
over time. The readjustment process can include two forms of inelastic
deformation;

 violent failures commonly referred to as rockbursts;


 slower creep-like movements.

Depending on the type of rock and the induced stresses, excavations


can show a propensity towards either of these two phenomena.

Creep is thus the continued deformation, or strain, of rock due to an


applied stress. Creep occurs on a time scale ranging from days to
years. The creep rate, denoted by ε‟, is a measure of the rate of change
of strain over a specified time, i.e. it is the derivative of strain with
respect to time;

d
 
dt
Creep rates are measured in strains or millistrains per second (m m-1 s-1
or mm m-1 s-1.

For the hard rocks of the South African gold mining industry, the steady-
state creep rate is typically of the order of 200x10-9 mm m-1 s-1 if the rock
is loaded to 90 % of its failure strength.

Some examples where the modelling of creep is required are:

 estimation of the long-term stability of service excavations,

Version 1.10:Jan 2010: UNSW: GS 9-3


MINING GEOMECHANICS

 determination of appropriate support and the optimum time when


support should be installed,
 estimation of the effect of mining rate on the volume of closure in
stopes, and
 estimation of the influence of mining rate on the stability of the
fracture zone ahead of advancing stope faces.

Although creep is a time-dependent characteristic of a rock mass, the


mathematical solutions employed in the modelling of creep are empirical
and not dynamic. This is because the modelling of creep-type rock
deformation does not involve wave propagation. Dynamic modelling, on
the other hand, is concerned with the physics of the situation and thus
does consider the propagation of waves through the material.

9.3 Modelling Creep Response of Rock


(Intact) Material
Some of the different kinds of creep phenomena that can be exhibited
by rock are exhibited by plotting strain against time t for which a sample
of rock is subjected to an applied stress.

The most common kind of creep response is represented by the curve


shown above. Following the loading strain ε0, the creep rate, as

Version 1.10:Jan 2010: UNSW: GS 9-4


MINING GEOMECHANICS

indicated by the slope of the curve, is high but decreases as the material
deforms during the primary (transient) creep stage.

At sufficiently large strains, the material creeps at a constant rate. This


is called the secondary or steady-state creep stage. Ordinarily this is the
most important stage of creep since the time to failure tf is determined by
the secondary creep rate .

The secondary creep stage is eventually interrupted by the onset of


tertiary creep, which is characterized by internal fracturing of the
material, creep acceleration, and finally failure.

A general equation for the total strain (εtotal(t)) experienced by a


specimen subjected to creep is:

 total (t )   0   p (t )  sst   T (t )

Where;

t is time

εo is the instantaneous elastic strain

εp(t) is a function expressing the decelerating creep of the


primary phase

ε‟ss is the rate of steady-state creep

εT(t) is a function expressing the accelerating creep of the tertiary


phase.

The steady-state creep rate is usually very temperature-dependent. At


low temperatures or applied stresses the time scale can be thousands of
years or longer. At high temperatures the entire creep process can
occur relatively rapidly. Another kind of creep response is shown by
unload curve. This is the sort of strain-time behaviour observed when
the applied stress is partially or completely removed in the course of
creep. This results in time-dependent or inelastic strain recovery.

Elastic creep and part of the transient creep curve may be recoverable if
the load is removed. The time dependent behaviour beyond this point is
termed “relaxation”. When time dependent shear behaviour occurs on
rock defects there may also be recoverable strain.

When a rock is tested, there is always hysteresis in the rock loading-


unloading stress-strain curve. The area under the curve represents
energy, so hysteresis indicates non-recoverable energy.

Version 1.10:Jan 2010: UNSW: GS 9-5


MINING GEOMECHANICS

Hysteresis in load-
displacement testing

Therefore generally we will consider only 3 stages of creep, ignoring


elastic strain.

The engineering models and laws that define creep can be subdivided
into three areas:

1. Empirical laws
2. Rheological models
3. Fundamental physical mechanisms

Empirical laws are developed from observation and may not necessarily
be based on fundamental material behaviour or property relationships.
Rheology is a general term for the time-dependent behaviour of

Version 1.10:Jan 2010: UNSW: GS 9-6


MINING GEOMECHANICS

materials. More simply, it is the study of flow and time-dependent


deformation of matter. Ultimately, an understanding of the underlying
fundamental mechanisms of creep will lead to the most robust
estimation and modelling techniques.

9.3.1 Empirical Creep Rate Law


Rocks behave like either;

 a brittle material (coalescence of microcracks or fracture from


cracks), or
 a ductile material (crack closure and plastic dislocation)

Which behaviour predominates depends on;

 Confining Pressure,
 Temperature,
 Cohesive Bond,
 Porosity, ...

The higher the cohesion and the lower the porosity, the more brittle is a
rock. However, ductile behaviour becomes more dominant with
confining pressure. Temperature and strain and loading rate have
influences on the behaviour. In general, stress in a deep natural rock is
strain-rate dependent:

 Q 
 
ss  A 1   3  e
n  RT 

Where;

A, n are material constants

σ1,σ3 are the major and minor principal stresses respectively

R is the universal gas constant (R = 8.31 J K-1mole-1)

T is the temperature of the sample in degrees Kelvin

Q is the creep activation energy. Q varies for the different creep


mechanisms and should be calculated from creep tests
performed over a range of temperatures.

Version 1.10:Jan 2010: UNSW: GS 9-7


MINING GEOMECHANICS

Alternately we can define the same compressive steady-state creep


in terms of strain:

ss  A1   3 n

where A and n are constants which be determined from fitting procedure


on graph of log (strain rate) versus log (stress), as will be seen later.

9.3.2 Rheological Creep Rate Models


It is possible to model creep associated with materials by assuming that
the material in question has viscous, or fluid, properties. It is then
possible to represent the time-dependent variation in strain, or strain
rate.

This is known as visco-elastic theory and is an extension of elasticity


theory, with the addition of one or more viscous components thereby
giving it a time-dependent nature. The models are formed from
assemblages of mechanical components, usually springs, dashpots and
sliders. They assist in understanding the material behaviour and allow
the formulation of the various constitutive relations.

In linear visco-elastic theory, various combinations of two principal states


of deformation;

 elastic behaviour (represented by a spring).


 viscous behaviour (represented by a dashpot).

are used to describe complex strain-time behaviour.

The spring, known as the Hookean model, represents elastic behaviour


and therefore gives an instantaneous strain (εo) that is constant with
time.

The dashpot, or Newtonian model, represents idealised viscous


behaviour and, after application of the load, gives a linearly increasing
strain with time. The slope of this curve is dependent on the viscosity
coefficient and the applied stress. A predefined viscosity coefficient,
denoted η, controls the dash pot response. The unit of the viscosity
coefficient is Pascal second (Pa s).

In the table below, some of the commonly used visco-elastic models are
shown. All are based on various combinations of the two basic
elements, Hookean and Newtonian models. The time-dependant strain

Version 1.10:Jan 2010: UNSW: GS 9-8


MINING GEOMECHANICS

profile shown alongside each model is based on a step-wise application


of a constant stress.

Burger's model, which is also known as the 4-parameter solid model, is


a combination of the Kelvin and Maxwell models in series. This model
provides an instantaneous response (εo), however, the strain does not
approach an asymptotic value but continues increasing linearly. When
comparing the typical creep profile associated with rock, it appears that
Burger's model provides the best representation of both the primary and
secondary phases of creep.

Version 1.10:Jan 2010: UNSW: GS 9-9


MINING GEOMECHANICS

If it is assumed that a laboratory rock sample under load follows


Burger's visco-elastic model, the time-dependent strain, ε(t), at a specific
instant in time is;

  
 EK t 
 t  2 (1 ) 

 (t )    1 e  K 

Em 21    m EK  
 

where EK and EM and ηK and ηM are the Young's moduli and viscosity
coefficients for the Kelvin and Maxwell units respectively, ν is Poisson's
ratio, and t is time.

The disadvantages of rheological models is that they:

 Provide no direct predictions unless calibrated. (The viscosity


coefficients of the models are not fundamental material
properties, but are dependent on the stress, temperature and
chemical environments. If a model is calibrated using the strain of
an experimental sample under load, the calibrated viscosity
parameters are only valid for the corresponding load).
 They do not account for shear and normal stress, temperature or
intrinsic structure
 Do not provide insight into fundamental creep mechanisms
 Linear visco-elastic theory cannot simulate the failure of rock.

On the positive side the advantages include:

 Allowing decomposition into elements


 Aid in engineering understanding, and
 Deficiencies can be overcome by
- Using various combinations of parallel and linear elements
- Making element parameters nonlinear

9.3.3 Fundamental Physical Mechanisms


In reality there is a blurring of the lines in distinguishing between
empirical laws and physical mechanisms. Use of term „physical
mechanisms‟ attempts to account for the strain dependence on stress
and temperature and derive relationships accordingly. We have already
seen that, in general, stress in a deep natural rock is strain-rate
dependent

Version 1.10:Jan 2010: UNSW: GS 9-10


MINING GEOMECHANICS

When a sample of rock is loaded at various strain rates (ε s-1), the log-
time behaviour can be determined. The stress function can also be
plotted as a stress difference (σ1 – σ3); and plotting log stress vs log
strain, n is slope of line.

Equations can be normalised so units are rational: f()/ 0 where o


is the shear modulus or initiating stress level.

 3 
Log (ss )  nLog  1   LogA
 0 

40
n
   
35 ss  A 1 3 
 0 
30

25 n=5
Log (εss)

20
n=6

15

10
n=3
log5A
n=1
0
0 2 4 6 8 10 12

Log [(σ1-σ3)/σ0]

These equations represent steady state creep (straight lines). They can
be used in combination to model other behaviour.

Laboratory Testing
Laboratory testing generally takes 3 forms, as illustrated on the next
page;

a. Constant stress state. A constant axial and radial stress is


applied and the strain is monitored. Constant stress creep tests
can be carried out in configurations such as triaxial, beam
bending, uniaxial, and direct tension.
b. Constant axial strain (constant strain rate). The strain rate
applied is constant and the stress is monitored. Constant strain
rate triaxial tests maintain a constant radial stress. Except in
rocks permitting steady state creep, this test will lead to yield
due to cumulative damage, and the results may be difficult to

Version 1.10:Jan 2010: UNSW: GS 9-11


MINING GEOMECHANICS

interpret as failure mechanisms may change as stress


increases.
c. Relaxation tests. A strain increment is applied and the stress is
allowed to relax. Relaxation tests require a stiff testing frame
and knowledge of the constitutive law form for a material in
order to interpret the results. Normally the test procedure is to
impose a instantaneous strain increment and then measure the
axial strain and stress decay with time.

Multistage triaxial tests are demonstrated here in the figures.

Version 1.10:Jan 2010: UNSW: GS 9-12


MINING GEOMECHANICS

Analysis of Laboratory Testing: Fitting Curves for Transient Behaviour


Either;

 Use log-log graphs as per rate process theory, OR


 Analyse time-strain graphs

Firstly identify steady state behaviour. Rule parallel line to steady-state,


beginning at the transient stage.

Separate plastic/elastic and steady state onto one graph and transient
behaviour on the other.

Fit a straight line equation to steady state behaviour;

 3 
Log (ss )  nLog  1   LogA
 0 

Version 1.10:Jan 2010: UNSW: GS 9-13


MINING GEOMECHANICS

and to transient (t) creep fit an exponential law similar in form to that of
a Kelvin model;

tr   t 1  e at  
Or

 0  
 E 
  t 
tr  1 e  EK  
E  
 

where EK is the Young's modulus for the Kelvin model and t is time.

9.4 Typical Creep Response of Rock (Intact)


Material
Hard Rocks
Low porosity rocks do not display significant TD behaviour under normal
conditions. Exceptions to this rule are when high stress regimes or
weathered material is encountered. The dominant mechanism is
microcrack generation & propagation along grain boundaries.

Coal
Time dependent components are related to mineral creep; gas
dissipation, strong fabric & weak pores. Pillar & high wall behaviour is
dependent on narrow clay (smectite) or thin coal seams. Coal pillars
progressively weaken due to internal fissuring and cleat fracturing. The
diagram shows how swelling clays can cause creep strain if defects are
dipping out of the slope (highwall) face, or cause a reduction in
confinement leading to pillar widening and extensional strain zones.

Version 1.10:Jan 2010: UNSW: GS 9-14


MINING GEOMECHANICS

Ultramafics
Ultramafics, especially those containing talcs and chlorites have a low
stiffness and thus begin to strain at lower stress magnitudes that stiffer
harder rock. They can also be subject to swelling.

9.5 Typical Creep Response of Rock


Discontinuities
Up to this point, all the models have been based on time dependent
(TD) behaviour of rock material. We will now briefly examine TD
behaviour of rock defects.

Earlier we mentioned that there are sometimes only 3 stages in


theoretical TD direct shear behaviour: primary, secondary and tertiary
creep. That is because shearing of defects involves movement of one
surface over another. If the surfaces are dry and there is no infill
present, then the movement is generally going to be permanent i.e.
there is no elastic component of the behaviour. Another key aspect is
that infill and moisture content have a large impact on behaviour.

A typical test result is shown below, showing creep stages as a function


of discontinuity shear displacement and time.

By examining typical triaxial tests of a discontinuity in hard rock, it is


found that;

 The instantaneous response (primary creep phase) of the rock


discontinuity exceeds the time dependent response (secondary
creep phase) by orders of magnitude.
 For all practical purposes the unfilled rock discontinuities are
inactive regardless of the amount of water present

Version 1.10:Jan 2010: UNSW: GS 9-15


MINING GEOMECHANICS

The shear creep rate of a joint is governed by the ratio of applied shear
stress to peak strength. The lower graph gives an example of this
behaviour. The ratio (horizontal axis) of shear stress/peak shear
strength varies from approximately 0.55 to 0.95 whilst the shear rate
varies from 0.00003 mm/hr to 0.00012 mm.hr representing a 3-fold
increase in shear creep rate.

Creep rate is also directly proportional to gouge thickness – the thicker


the gouge the higher the creep rate, and moisture content has a
significant effect on creep rate: higher creep rates and larger
displacements with increasing moisture content.

For a discontinuity, the following model was proposed for secondary


or steady state creep:

Version 1.10:Jan 2010: UNSW: GS 9-16


MINING GEOMECHANICS

n
 
D ss  A  
 
 p

where

Dss = Shear deformation rate


 = Applied shear stress
p = Peak (Coulomb ) shear stress
A&n = constants obtained from a log-log graph fit of test
data (Log(creep rate) against Log(shear stress/shear
strength).

Obtaining the coefficient value for n is obtained from the slope of the line
on log-log graph. The coefficient Log A is obtained from the intercept.

In the example below:

n= 3.48; and log A= -10.45

Version 1.10:Jan 2010: UNSW: GS 9-17


MINING GEOMECHANICS

9.5 Dynamic Behaviour of Rock

9.5.1 Introduction
Dynamic behaviour of rock describes the rock response to the rapid
strain rate induced by explosives and/or mine seismicity. When the rock
mass is subjected to a sudden change of the equilibrium conditions the
particle closest to the disturbance is moved from its state of equilibrium
and affects the neighbouring particles; the disturbance propagates
through the rock mass and transports energy but no material.

When considering rock testing, there are orders of magnitude difference


between seismic strain and laboratory strain - upto 10 orders of
magnitude. An increase in strain rate generally leads to an increase in
Young‟s modulus, an increase in UCS (for a given strain) and an
increase in the brittleness of the rock behaviour.

The re-distribution of stresses caused by a sudden removal of rock


during tunnelling and/or failure of a part of the rock mass leading to a
decrease in the stiffness are examples where strain waves are
generated and then propagated through the rock mass as illustrated
below.

Another type of disturbance is generated by the detonation of explosives


and the occurrence of rock bursts and earth quakes. A third type of
disturbance applied to the rock mass is vibrations from, for instance,
drilling machines and trains. The disturbances of the first and the
second category generate transient strain waves with a short duration
whereas the third category is more or less periodic. The influence of the
dynamic events in the rock mass can be analysed in two ways:

(i) the effect of reflection and refraction of transient waves

Version 1.10:Jan 2010: UNSW: GS 9-18


MINING GEOMECHANICS

(ii) analysis of the vibrations of the rock mass caused by the


event.

If the main sources of dynamic behaviour (as outlined above) are:

 Failed rock mass


 Explosives
 Machinery

Where do they fit into the general rock engineering process?.

Site Characterisation (Data collection and Analysis)

Mine model formulation (Engineering Analysis)

Design (Global & local)

Implementation

Rock Mass performance monitoring

Retrospective Analysis

Dynamic behaviour caused by failed rock masses or movements along


faults will be detected via microseismic monitoring. Dynamic behaviour
of rock types should be taken into account with drill and blast design;
which can be considered as an implementation issue (cf Design refers
to excavation size/shape). The drill and blast design can also be
optimised via monitoring and analysis of fragment distribution.
Laboratory and field tests can assist in determining the dynamic
properties of rock or rock masses. Finally, the dynamic behaviour due
to machinery is generally low impact and thus ignored.

9.5.2 Seismic Waves and Noise


Stress waves are the expression of dynamic stress changes and are
essentially sound waves in solid material. Two types of stress wave are
propagated within the rock mass:

 one has particle motion in the x-direction (longitudinal Primary or


P-waves),

Version 1.10:Jan 2010: UNSW: GS 9-19


MINING GEOMECHANICS

 one has particle motion in the y- or z-directions (transverse,


Secondary or S-waves).

Two other types of stress wave which are important are Rayleigh and
Love waves. Both of these waves occur near interfaces and free
surfaces and have elliptical particle motion which is polarized and
perpendicular to the free surface:

 Rayleigh waves: motion is parallel to the direction of propagation


 Love waves: motion is perpendicular to the direction of
propagation

Rayleigh waves are the most destructive wave, as they occur during
earthquakes. In general, if a P-wave reaches a boundary, which is not
parallel to the wave front, four waves are generated. Two of these are
reflected waves, moving back into the medium from which the original
wave came, a P-wave and a S-wave; the other two waves, also a P-
wave and a S-wave, are transmitted into the new medium.

On the next page is an example of the waveforms from a mining


induced seismic event in South Africa recorded at five different local
sites using geophones. The horizontal axis represents time and the
vertical axis represents velocity. This demonstrates how the stress
waves move with different velocities: the p-wave is always the first to
arrive.

Version 1.10:Jan 2010: UNSW: GS 9-20


MINING GEOMECHANICS

Below is an enlarged view of the 4th waveform on the above seismic


trace.

What you see is noise - an unwanted signal. Seismic noise is seismic


energy other than the principal arrivals or reflections and is sometimes
referred to as ambient seismic noise due to background earth
movements. The amplitude of seismic noise is generally very small as
demonstrated here. Noise is temporally and spatially variable and tends
not to be uniform at all frequencies. It is possible to remove noise via
filtering techniques.

9.5.3 Wave Velocity and Measurement


We refer to the velocity of the p- and s-waves as Vp and Vs respectively.
The velocity of each of these waves depends on the density and elastic
properties of the rock mass. They can be calculated via the following
expressions:
1
E (1   )  2
Vp   
  (1   )(1  2 ) 

1
E 1  2
Vs   
  2(1   ) 

1
Vp  2(1   )  2

Vs  1  2 

Where

Version 1.10:Jan 2010: UNSW: GS 9-21


MINING GEOMECHANICS

Vp is the P-wave velocity


Vs is the S-wave velocity
 is the density of the rock/rock mass
 is Poisson‟s ratio
E is the modulus of elasticity

The ratio of Vp to Vs depends on the Poisson‟s ratio and is usually


between 1.4 and 1.9 The plot below demonstrates this relationship with
the ratio Vp/Vs on the vertical axis and Poisson‟s ratio on the horizontal
axis.

3.5

2.5

2
Vp/Vs

1.5

0.5

0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45

We need to have estimates of the dynamic elastic properties of rock


material to make assumptions about a rock mass. To this end we can
perform laboratory tests to ascertain values for dynamic elastic and bulk
modulus and Poisson‟s ratio. This measurement is performed on core
samples prepared for UCS testing. The velocities of compressive and
shear ultrasonic waves through the core sample are measured and
used to calculate the elastic modulus and Poisson's ratio. This method
indicates the competency of the rock. Other factors such as anisotropy
affect the results, however, and a minimum of five measurements is
recommended.

From the waves‟ velocities and the sample bulk density, the dynamic
elastic modulus and dynamic Poisson‟s ratio are then calculated.

Dynamic Elastic Properties


We have already reviewed the fact that faster strain rates result in
increased elastic modulus. It therefore follows that the dynamic moduli
are higher than the static moduli. On average the increase is around
30% but it can be up to 200% (or a ratio of 3:1).

Version 1.10:Jan 2010: UNSW: GS 9-22


MINING GEOMECHANICS

The dynamic values can be determined from the measurement of Vp


and Vs such that;

3VP2VS2  4VS4
E
VP2  VS2

VP2  2VS2

2(VP2  VS2 )

G  VS2

where

Vp is the P-wave velocity


Vs is the S-wave velocity
 is the density of the rock/rock mass
 is the dynamic Poisson‟s ratio
E is the dynamic modulus of elasticity

G is the dynamic bulk modulus

It is important to note that small errors in Vp and Vs can lead to large


errors in Poisson‟s ratio.

Impedance Matching
Explosives with high brisance (or high shock energy) are well suited for
blasting in hard competent rock. A low brisance explosive such as
ANFO (which produces a lower shock wave but relative high gas or
heave loading) is better suited for soft porous, or heavily fractured rock.
The best matching for optimum shock wave to the rock occurs when the
detonation impedance D equals the impedance  of the rock material,
Vp.

Although emulsions appear to approximate match sandstone and some


types of granite, in terms of impedance, „designer‟ impedance matching
is not achieved, as explosive costs would make this uneconomic.

Version 1.10:Jan 2010: UNSW: GS 9-23


MINING GEOMECHANICS

Impedance μ explosive
Detonation Velocity, D

P-wave velocity, Vp

Density ρ (kg/m )
3

-2 -1
10 kgm s
or rock
(m/s)

(m/s)

6
Explosive

ANFO (d=125mm) 3200 900 2.9

Emulsion 5200 1200 6.2

Rock Material

Basalt, dense 5560 2761 15.4

Granite 2710 2640 7.2

Granite, dense 5230 2800 14.6

Hematite ore 6280 5070 31.8

Sandstone 2640 2182 5.8

Concrete 4580 2220 10.2

Stress and Strain Generated By Wave


As a shock wave is transmitted into the (jointed rock mass) medium
surrounding the charge in a drill hole, the wave energy or the shock
wave amplitude, which is identified by the peak particle velocity PPV,
decreases with distance from the hole. When a stress wave σ1 travels
at a sound speed c1 in a medium with density ρ1, its impedance (μ1) is
the product of the density and the sound wave velocity;

1  1c1

At a given distance (or time) identified by the peak particle velocity


PPV1, the elastic stress (σ1) and strain (ε1) generated by a P-wave is
given by;

1 PPV1
 1  1 PPV1 1  
E1 c1

Version 1.10:Jan 2010: UNSW: GS 9-24


MINING GEOMECHANICS

References
Brady B.H.G, & Brown, E.T., 1999, Rock Mechanics for Underground Mining 2nd.
Ed. Chapman and Hall, Oxford
Hudson J A & Harrison J P (1997) Engineering Rock Mechanics, An Introduction to
the principles, Pergamon.
Potvin Y, & Nedin, P, 2003, Management of Rockfall Risks in Underground
Metalliferous Mines, A reference manual, Minerals Council of Australia,
Kingston
Weijermars, R. 1997. Principles of Rock Mechanics, Alboran Science Publishing,
Amsterdam
Malan, D.F. (1999) Time-dependent behaviour of deep level tabular excavations in
hard rock, Rock Mech. Rock Engng, Springer-Verlag, Austria, Vol. 32, No.2,
pp. 123-155. Pande, G.N., Beer, G. and Williams, J.R. (1990) Numerical
Methods in Rock Mechanics, John Wiley & Sons Ltd, Chichester.
Dusseault, M.B. and Fordham, C.J. (1993) Time-dependent behaviour of rocks, In:
Comprehensive Rock Engineering, Vol. 3 (edited by JA Hudson), Pergamon
Press, Oxford, England, pp. 119-149.

Version 1.10:Jan 2010: UNSW: GS 9-25


Owner List
The Owners-SP 2657

Lot Unit Name and Main Contact Address Levy Address Notice Address Entitlement Balance
(-)prepaid
1 1 SAMIR BAHIG SHOUSHA 1 -1.46
2/62 MASCOT DRIVE, EASTLAKES NSW 2018 2/62 MASCOT DRIVE, EASTLAKES NSW 2018 2/62 MASCOT DRIVE, EASTLAKES NSW 2018
2 2 MR S B SHOUSHA 1 -2.45
2/62 MASCOT DRIVE, EASTLAKES NSW 2018 2/62 MASCOT DRIVE, EASTLAKES NSW 2018 2/62 MASCOT DRIVE, EASTLAKES NSW 2018
3 3 EDWARD ADHIKARI & DAISY BISWAS 1 0.00
3/62 MASCOT DRIVE, EASTLAKES NSW 2018 3/62 MASCOT DRIVE, EASTLAKES NSW 2018 3/62 MASCOT DRIVE, EASTLAKES NSW 2018
4 4 SEAN & CATHERINE O'DONOGHOE 1 0.00
4/62 MASCOT DRIVE, EASTLAKES NSW 2018 4/62 MASCOT DRIVE, EASTLAKES NSW 2018 4/62 MASCOT DRIVE, EASTLAKES NSW 2018
5 5 MR DAVID HENRY POPPLEWELL 1 0.00
5/62 MASCOT DRIVE, EASTLAKES NSW 2018 5/62 MASCOT DRIVE, EASTLAKES NSW 2018 5/62 MASCOT DRIVE, EASTLAKES NSW 2018
6 6 MR G & MRS U KASSAR 1 0.00
2 KYOGLE STREET, EASTLAKES NSW 2018 2 KYOGLE STREET, EASTLAKES NSW 2018 2 KYOGLE STREET, EASTLAKES NSW 2018
7 7 MR J & MRS M REPETTO 1 0.00
42 HIGHCLERE AVENUE, ROCKDALE NSW 42 HIGHCLERE AVENUE, ROCKDALE NSW 42 HIGHCLERE AVENUE, ROCKDALE NSW
2216 2216 2216
8 8 TRAM THI NGOG LE 1 0.00
8/62 MASCOT DRIVE, EASTLAKES NSW 2018 8/62 MASCOT DRIVE, EASTLAKES NSW 2018 8/62 MASCOT DRIVE, EASTLAKES NSW 2018
9 9 MR & MRS MAHMOUD 1 404.80
9/62 MASCOT DRIVE, EASTLAKES NSW 2018 9/62 MASCOT DRIVE, EASTLAKES NSW 2018 9/62 MASCOT DRIVE, EASTLAKES NSW 2018
10 10 YOHANES SOE & HANG LIN WIE 1 404.80
10/62 MASCOT DRIVE, EASTLAKES NSW 10/62 MASCOT DRIVE, EASTLAKES NSW 10/62 MASCOT DRIVE, EASTLAKES NSW
2018 2018 2018
11 11 MESSRS T, A & J & MRS K JAKUS 1 0.00
11/62 MASCOT DRIVE, EASTLAKES NSW 11/62 MASCOT DRIVE, EASTLAKES NSW 11/62 MASCOT DRIVE, EASTLAKES NSW
2018 2018 2018
12 12 MRS Y KASSAR 1 0.00
68 FLORENCE AVENUE, EASTLAKES NSW 68 FLORENCE AVENUE, EASTLAKES NSW 68 FLORENCE AVENUE, EASTLAKES NSW
2018 2018 2018

28/08/2009 18:49 Mark Gitman MG Strata and BMC Management Pty Ltd Page 1
Rock
MINING GEOMECHANICS

Mechanics

10
TOPIC 10: SLOPE
STABILITY

10.1 Introduction
A fairly common engineering failure is that of embankments, cuttings or
slopes (natural or man-made). The causes of these failures are very
often water-related, either through shear strength reduction, and /or
additional lateral loading as a result of tension-cracks, etc.

Slope instability in rock is often related to either;

 discontinuity controls (i.e. failure of the slope along joints, faults,


etc.)
 the effective stress state in the slope (i.e. a very weak and/or
highly fractured rock mass can behave like a soil). In these
circumstances the slope fails along a surface representing the
path of least resistance through the material.

Anisotropic material properties and major structural features can


produce a failure profile that deviates from the ideal circular geometry.
Heavily broken rock tends to behave like a cohesionless soil, producing
a failure profile that is almost planar (i.e. has a very large radius of
curvature) dipping at an angle close to the angle of friction of the broken
rock. Although the failure surface usually passes through, or very close
to, the toe of the slope, variations in material strength or water conditions
can produce limited zones of slope failure.

Stope stability analysis methods can be used to analyse a failure that


has already occurred. In these back analyses the geometry of the slip
surface profile is usually known with a fair degree of accuracy; the aim is
to estimate reliable values for cohesion c' and the angle of friction Φ' for
the failed mass, for use in future design. The reliability of the estimates

Version 1.10:Jan 2010: UNSW: GS 10-1


MINING GEOMECHANICS

for c' and Φ’ is usually critically dependent on the accuracy of


information concerning water conditions in the slope immediately prior to
failure.

In most cases there will not be existing slope failures to analyse, so the
shear strength parameters must be determined from appropriate
laboratory tests, field tests and on-site measurements. The geometry of
the failure profile will be unknown, so it is necessary to identify the failure
profile that gives the very lowest factor of safety. This profile, which will
dictate the stability of the slope, is called the critical failure surface.

10.1.1 Principal Factors Affecting Slope Stability


The steps in the analysis of slopes in general are;

Assess the loading conditions: A cutting is an unloading


case where soil is removed, causing relief of stress
in the rock or soil and consequent change of shear
strength. Resistance dissipates over time and
shear resistance may well reduce in the long-term.
Embankments, spoil heaps, etc. are on the other
hand a loading case and the construction period is
the most critical, owing to the build-up of pore
pressures and the reduction in effective stress. In
the long-term, pore pressures dissipate and shear
resistance increases.

Determine the available shear strength in the vicinity of


the sliding surface (or discontinuity planes on which
rock blocks can slide) and evaluate the lower bound
(most critical or limiting) conditions. Take the true
factor of safety to be the lowest of those values.
Several evaluations are often required to locate the
‘critical’ or lowest FOS slip surface.

Assess the water flow conditions in the media in order to


determine the water pressures and consequently
the water thrusts and the uplifting forces acting on
the potential unstable mass. Also determines
whether total or effective stress and strength
parameters are used.

The stability analysis is then carried out for the


potentially unstable mass in order to determine the
slope safety factor or to determine the motion of the
blocks detached from the rock slope for rock fall
analysis.

Version 1.10:Jan 2010: UNSW: GS 10-2


MINING GEOMECHANICS

Finally, the efficiency of stabilizing technique applications


(excavation, drainage, rock bolts, soil nails, etc.) are
computed still using the stability analysis for the
cases in which the slope stability conditions need to
be improved.
For a rock slope in which instability is discontinuity controlled (plane,
wedge failure, etc.), step 1 is prefaced by the analysis of the
discontinuity sets and the potential instability kinematics.

10.2 Limiting Equilibrium

10.2.1 Factor of Safety


In order to compare the stability of slopes, or the comparative stability of
a single slope with various failure geometries, some form of index is
required and the most commonly used index is the Factor of Safety
(FOS). This can be defined as;

the ratio of the total force available to resist sliding to the total force
tending to induce (mobilise) sliding.

The factor of safety (FOS) of a slope is defined as;

 Forces resisting failure


FOS 
 Forces mobilising failure

The condition of limiting equilibrium is defined as the condition of all


unstable mass forces being in balance (i.e. the mass is on the point of
failure). In this case the FOS = 1,00.

As an example, consider a block resting on a slope of angle  to the


horizontal, with coefficient of base friction :

N

Forces parallel to slope are thus;

S  W sin 

And forces normal to slope are

Version 1.10:Jan 2010: UNSW: GS 10-3


MINING GEOMECHANICS

N  W cos 

Thus;

S W sin
  tan 
N W cos

If  > tan then we say that not all of the frictional resistance to sliding is
mobilised. If  = tan, we have a state of limiting equilibrium.

The factor of safety against sliding would thus be;


FOS 
tan 

so factor of safety is defined as the factor by which  must be divided to


attain a state of limiting equilibrium.

Where there is both friction and cohesion (coefficient of friction tanΦ’


and cohesion c’), the factor of safety is defined as the factor by which
both tanΦ ‘and c’ must be divided to bring the slope to a state of limiting
equilibrium. Whenever the forces resisting shearing failure are larger
than those causing it, then the slope will be stable. When designing or
analysing a slope a minimum FOS is required, which depends on the
type of excavation (mining or civil) and a number of other associated
factors. In essence, consider the risk associated with the slope failure
and the results of failure in terms of;

 Life of the slope - short or long term


 Type of slope - bench, stack or overall slope
 Mine or civil slope - production or final slope
 Location of installations above or below slope
 Exposure of people and equipment to slope

The table on the next page summarises some FOS guidelines.

Especially in mining, a flatter slope (and often therefore more stable) will
require a larger stripping ratio to mine it and thus profitability of the mine
is adversely influenced. In the end, a cost-benefit-risk analysis can be
undertaken to investigate the cost penalties associated with flatter
slopes, compared to the cost penalties associated with clearing-up
failed. In using this approach - you are effectively putting a finite value
on people’s lives (mine worker - the public, etc.) and this may not be
ethically defensible.

Version 1.10:Jan 2010: UNSW: GS 10-4


MINING GEOMECHANICS

Example Consequence of Minimum


failure acceptable
FOS

Individual benches Not serious

Small temporary slopes not Low public exposure


adjacent to roads or buildings 1,3
Easy to remove failed
Bench stacks not adjacent to mass, non disruptive
roads or buildings, short term
only

Any benches or stacks of a Moderately serious


permanent or semi-
permanent nature. Low public exposure
1,5
Benches or stacks above or Disruptive to clear-up
below a building or haul
roads

Major stacks or overall Serious


slopes
Medium public
1,8
Stacks or benches carrying exposure
main haul roads or above or
below permanent Extremely disruptive
installations clear-up

Any long term civil Very serious - law suits


excavation and damages would
>2,0
apply

High public exposure

For example, a FOS of 1,3 (or 30% more resistive than mobilising shear
force), would be typical for most short term mining benches. It must be
remembered that depending on the sensitivity of the slope AND the type
of analysis conducted, some destabilising factors may not be included in
the simple limit equilibrium approach and therefore it is good practice to
adjust the FOS required, bearing in mind the quality of the data used in
the analysis and how rigorous the analytical procedure is.

Although instability is often analysed in terms of a limiting equilibrium


factor of safety approach only, instability risk is a concept which should
be introduced into stability analysis because of the uncertainties
connected to the geological material parameter determinations.
Probabilistic analyses are carried out for this purpose. Geostatistical
models for the geological variables governing slope stability problems

Version 1.10:Jan 2010: UNSW: GS 10-5


MINING GEOMECHANICS

can also be carried out, by considering the regionalized aspect of those


variables and by therefore decreasing the uncertainties of the slope
stability assessment. These approaches help to define the real
variability of a factor of safety, based on the natural variability of the field
data used in the analysis.

10.3 Modes of Failure


In this Module we consider only sliding material instabilities. Sliding
movements are further divided into rotational and translational slides.

A sliding movement is determined by unbalanced shear stress along


one or more (slip or failure) surfaces. These surfaces are visible or may
be inferred by analysing 'in situ' conditions or the worst case scenario
(associated with lowest factor of safety). Sliding surface determination is
one of the most important problems in a slope stability analysis. Two
basic situations exist with respect to sliding surface location;

 A failure may already have occurred and the sliding surface


shape and position can be identified by means of site
investigations. In this case slope stability analysis is carried out
with the purpose of assessing unstable slope mass strength or
pore pressure parameters by means of a back analysis (but
limited in terms of the normal force and instantaneous c’ and Φ’
values that applied at the time of failure).
 An instability can be incipient or a sliding condition can be
determined for a man-made slope. In these cases, slope stability
analyses are carried out with the purpose of locating a critical
failure surface and associated factor of safety. The methods
proposed to automatically locate the critical failure of a slope are
by considering it to be the surface that has the minimum safety
factor. This procedure of minimum safety factor determination
has disadvantages and limitations.
The three types of sliding movements considered here are shown
below;

Version 1.10:Jan 2010: UNSW: GS 10-6


MINING GEOMECHANICS

CIRCULAR SLIDES
Observed failures in relatively homogeneous or heavily fractured and
weak material often occur along curved failure surfaces. A circular slip
surface, like that shown in the Figure above (a), is often used because it
is convenient to sum moments about the center of the circle, and
because using a circle simplifies the calculations. A circular slip surface
must be used in the Simplified Bishop Method.

PLANAR (WEDGE) SLIDES


Planar slides are characterized by slides on a single plane (so called
‘infinite slope’ instability) or two intersecting planes (multiple wedge
failure, similar to that found with retaining wall active failure). Slides are,
in this case, translation movements. Much of the material is forced
down the slide plane as masses of ‘intact’ material with failure on certain
defined planes or failure surfaces.

Wedge failure mechanisms are defined by three straight line segments


defining an active wedge, central block, and passive wedge as shown in
the Figure (b). This type of slip surface may be appropriate for slopes
where the critical potential slip surface includes a relatively long linear
segment through a weak material bounded by stronger material. A

Version 1.10:Jan 2010: UNSW: GS 10-7


MINING GEOMECHANICS

common example is a relatively strong rock dump or embankment


founded on weaker, stratified alluvial soils.

GENERAL NON-CIRCULAR ROTATIONAL SLIDES


The most common rotational sliding phenomena involve a sliding
surface with a spoon shape or a convex cylindrical shape. The sliding
surface is seldom a uniform concave upward spherical segment; often
the presence of soil horizons and non-homogeneity zones influence the
sliding surface shape. The sliding surface is therefore often composed
of straight lines and circular arcs. It is especially important to investigate
stability with noncircular slip surfaces when soil shear strengths are
anisotropic. The diagram below summarises the influence of material
type on slip surface development.

Version 1.10:Jan 2010: UNSW: GS 10-8


MINING GEOMECHANICS

To summaries, the modes of failure discussed above imply failure due to


sliding of one or more masses of material on slip surfaces, without
plastic strain developing within each of these masses, as if they were
rigid bodies. This is exactly what would be expected of a strain
softening material. It is assumed in this analysis that if masses of soil
move as rigid bodies, they must do so in a kinematically feasible
manner, i.e. there must be no overlaps of or gaps between adjacent
masses of soil.

10.4 Stability of Cohesive-Frictional Materials

10.4.1 Slice-based Slip Circle Analysis


Section 10.4.1 has been adapted, with minor modifications, from the Soil
Mechanics Module (Watson, 2007) and Slope Stability in Heavily
Fractured Weak Rock (Priest, 2010), and are reproduced here with
permission.

The approach considers the slope sliding mass to be comprised of a


number of slices, each of which contributes to the failure, either through
downslope forces exceeding resistive forces, or (especially for deep
seated failures) resistive forces exceeding downslope forces.

The shape of the known, or assumed, failure profile can be divided into
a number of distinct linear elements. There will be three unknown
values for each element:

 the reactive shear force


 the reactive normal force
 the location of the reactive normal force.

For a two-dimensional planar analysis there will be three equations of


static equilibrium for the potentially unstable mass:

 the sum of the vertical forces,


 the sum of the horizontal forces
 the sum of the rotational moments.

Clearly if the mass has been divided into more than one element there
will be more unknowns than there are equations, and the problem will be
statically indeterminate. To make the problem soluble it is necessary to
generate more equations to solve for the unknowns. These equations
can be generated by dividing the mass into a number of slices, which

Version 1.10:Jan 2010: UNSW: GS 10-9


MINING GEOMECHANICS

are usually vertical and, to simplify calculations, are of equal width, as


shown below;

R
i

slice 1

slice i

bi
slice n

A minimum of 8 are generally required; increasing the number of slices


beyond 20 provides a negligible improvement in accuracy.

The principles of static equilibrium can be applied to determine slope


stability by the method of slices. The general slope geometry is
summarised below. The slope consists of a crest and toe, with a circular
or multi-linear failure profile passing through the toe. A vertical, or near
vertical, tension crack may connect the upper part of the slip surface
with the top of the slope.

Top of slope
Crest of slope
Tension crack
Typical slice

Toe of slope
Slip surface

The main assumptions are as follows:

1. The shear strength of the material at the base of each slice in the
slope is governed by the Coulomb or Hoek-Brown yield criterion.
If the Hoek-Brown criterion is adopted then it will be necessary to
calculate the equivalent instantaneous values of cohesion and
angle of friction c'i and 'i for the base of each slice (refer to Topic
5). Each slice can have different shear strength parameters c'
and tan '. The ‘dash’ means that the values of cohesion and
friction are the effective strength parameters, recognising the
important influence of water pressure in controlling slope failure.

Version 1.10:Jan 2010: UNSW: GS 10-10


MINING GEOMECHANICS

2. It is implicit in this assumption that all of the slices have the same
factor of safety, even though their shear strength parameters
may be different.
3. The slices are long in the direction normal to the plane of the
cross-section and the forces on the ‘out-of-plane’ boundaries of
each slice are negligible.
4. The normal and shear reactions on the left side and right side of
each slice are assumed to cancel each other out.
5. The profile of the slope surface, the phreatic (water) surface and
the location of the tension crack (if any) are known.

Slice width b Slope surface

Water table

hw hr ER
XL
XR
EL
Typical slice
W
S

N’

For a slice of unit breadth normal to the plane of the cross-section, the
following parameters are defined:

 All slices are assumed to have the same width b. The height of
rock (or soil) in a typical slice is hr and the vertical height of the
water table above the base of the slice is hw
 The unit weights of rock (or soil) and water are respectively r and
w.
 The normal and shear reactions on the base, left side and right
side of the slice are respectively N, S, EL, XL and ER, XR, as
shown in the diagram. The reactions EL, XL and ER, XR are
assumed to cancel each other out, and will be ignored.
 The inclination of the base of the slice is . The diagram shows
the positive sense for ; a slice with a base that slopes
downwards from left to right (the mirror image of the slice shown
above)would have a negative value for .

Version 1.10:Jan 2010: UNSW: GS 10-11


MINING GEOMECHANICS

The weight W of a typical slice is given by

W  hb r

The area A of the shear surface at the base of the slice is given by

b
A
cos

The water pressure u on the base of the slice is given by

u  hw w

Under conditions of limiting equilibrium, the sum of the vertical forces


acting on each slice will be zero. Taking downward acting forces to be
positive, the vertical components of the forces acting on the slice are

W  S sin   N cos   0

The effective normal force N' on the base of the slice is given by

N   N  uA

From which the following expression is developed;

W  ub
N   S tan 
cos

Substituting for A and N' in the Coulomb criterion, the shear strength (as
a force)is;

S f  N  tan    cA

So

cb  W  ub 
Sf    S tan   tan  
cos  cos 

At limiting equilibrium, S = Sf , hence with Factor of safety (FOS) of 1.00


then at the base of the slice;

cb  W  ub  tan   hence


S    S tan  
FOS cos  cos  FOS

Version 1.10:Jan 2010: UNSW: GS 10-12


MINING GEOMECHANICS

 tan  tan    cb  W  ub  tan  


S 1     and
 FOS  FOS cos  cos  FOS

S
c b  W  ub  tan  sec 
 tan  tan   
FOS 1  
 FOS 

The shear reaction S on the base of the slice, and the factor of safety of
the slope FOS are unknown. In order to determine the value of S it is
necessary to make further assumptions about how the rock/soil mass
moves. The simplest of these mechanisms is to assume that the mass
fails by rotation along the arc of a circular slip surface of radius R, as
shown below.

Consider the moment equilibrium of the entire sliding mass of soil


consisting of n slices.

Centre of slip
circle of radius R R z

H = horizontal
R hydraulic thrust
in tension crack
W

S
 N'

IF a tension crack is included at the top of the slope (which is typical for
this failure geometry), moments are;

 the horizontal thrust H in the tension crack exerting a moment z


about the centre of the slip circle;
 the weight W of a slice exerts a moment x (x = R sin 
 the shear reaction S exerts a moment R.

Note N' acts radially so it exerts zero moment.

Taking clockwise moments about the centre of the circular slip to be


positive, at limiting equilibrium;

Version 1.10:Jan 2010: UNSW: GS 10-13


MINING GEOMECHANICS

n n

Wi R sin i  Hz   Si R
i 1 i 1

Dividing by R and substituting for S;


n
Hz n
cb  W  ub tan  sec
 Wi sin  i  R

 tan  tan   
i 1 i 1
FOS 1  
 FOS 

Multiplying by FOS and dividing by the LHS gives:

 
 
1 n
 cb  tan   W  ub
Hz 
FOS  n   
  tan   
 W sin  
R
i 1

 
1
tan
 
i 1
  F  

Since FOS appears on both sides of the equation, it is necessary to


compute FOS iteratively, starting from an initial guess FOSJ-1 to
determine FOSJ. The value for FOSJ is then substituted for FOSJ-1 and
a new value determined for FOSJ = 1, 2, … until the charge in FOS is
negligible. Convergence is usually achieved after three or four
iterations.

So the solution for i to n slices becomes with iterative estimates of FOSJ


is;

 
 
1 n
 cibi  tan i Wi  ui bi 
Hz 
FOS j  n   
tan i tan  i  
 Wi sin  i  i 1

 
1  
 
i 1 R F j 1

Sources of error
 Many circles must be considered to ensure the slip circle with the
lowest (critical) FOS is found – this is where failure will occur.
 Horizontal and moment equilibrium of a slice are not assured,
and the interslice forces are ignored in vertical equilibrium.
 No allowance is made for resistance to slip at the ends of the
slide (the slip surface is spoon shaped when considered in 3-D).
 Soil is not an elastic – perfectly plastic material, so peak strength
is not mobilised simultaneously at every point of the slip surface.
 It is supposed that the given c’ and tan Φ’ will be operative
simultaneously at all points of the slip surface. In fact the failure

Version 1.10:Jan 2010: UNSW: GS 10-14


MINING GEOMECHANICS

is progressive, beginning at the top and bottom of the slip surface


and eventually meeting in the middle, by which time soil strength
over much of the surface has fallen to residual.

10.4.2 Non-circular Slip Surfaces


If the soil mass is not of uniform strength, the lowest FOS may be
associated with a slip surface running through the weaker material. The
sliding mass no longer moves as a rigid body, so it is instead considered
as a number of wedges or slices which slide with respect to each other.
In circular failure, work was being done on the shear surface only, in
non-circular failure, work is being done on the slip surface AND on the
planes of contact between the slices, hence interslice forces (Z) should
be considered.

There are several slice-based analyses that cater for non-circular slip
surfaces, of them, the Morgenstern-Price analysis is the most accurate
for general (non-circular) slip surfaces. The Spencer or Spencer-Wright
analysis (e.g. GALENA) is essentially the same as Morgenstern-Price.
The Sarma and multiple wedge analyses are simpler versions of
Morgenstern-Price, probably less accurate.

A further reduction of the Sarma and multiple wedge analysis leads to a


two- or three-wedge analysis in which the slip surface is approximated
by wedge-shaped slices.

Multiple Wedge Analysis


Core dams (an earth dam with an impermeable clay core) are examples
of where a slope failure can be analysed using a multiple wedge
analysis. Wedges are often referred to as ‘active’ zones (tend to drive
the instability), ‘passive’ zones (tend to resist the instability) and (if three
or more wedges analysed), ‘central’ zones.

The inclination of the forces on the vertical boundaries between the


zones are assumed (the interslice force between the central block and
the passive wedge is sometimes assumed to be horizontal). The
Wedge Method is actually a special case of the force equilibrium
procedure and fully satisfies equilibrium of forces in the vertical and
horizontal directions and ignores moment equilibrium.

In the two-wedge active-passive wedge problem shown on the next


page, (with no water pressure considered), the forces between the
wedges are usually taken to act at an angle α which, following the
Morgenstern-Price approach, is taken to be the angle of friction of the
sliding mass.

Version 1.10:Jan 2010: UNSW: GS 10-15


MINING GEOMECHANICS

NOTE that in the two wedge case shown below, it can be expected that
the friction angle of the sliding mass and the core (on which the active
wedge slides) are NOT the same.

Active
wedge

Passive
wedge

S1

S1 N1
Z1
W1
N1
Z2
W1
S2 Z1=Z2 N2
W2 N2
W2

S2

In the special case of a wedge boundary normal to the active slip


surface, the angle α is often taken parallel with this slip surface (see Zp
in the three wedge problem below).

Version 1.10:Jan 2010: UNSW: GS 10-16


MINING GEOMECHANICS

The most straightforward method of solution is to use equations of static


equilibrium and substitute for N and S in terms of W and Z, with c and
tanΦ divided by FOS. The two expressions can be re-arranged as
quadratic equations in Z1, Z2. When a three wedge problem is
encountered, a cubic equation is obtained instead.

A polygon of forces approach can also be used to resolve Z1, Z2. or Za


and Zp.as appropriate.

References
Priest S. D. (2010). Slope Stability in Heavily Fractured Weak Rock. Lecture
Notes, Mining Geomechanics. The University of Adelaide.
Watson, J (2007). Soil Slope Stability. Lecture Notes, Mining Geomechanics,
University of New South Wales.

Version 1.10:Jan 2010: UNSW: GS 10-17

You might also like