Sunteți pe pagina 1din 213

Engineering

Behaviour of
Rocks
Engineering
Behaviour of
Rocks
Second Edition

IAN FARMER

LONDON NEW YORK


CHAPMAN AND HALL
First published 1968 by
E. & F. N. Spon Ltd
Second edition 1983
Published by Chapman and Hall Ltd
11 New Fetter Lane, London EC4P 4EE
Published in the USA by
Chapman and Hall
733 Third Avenue, New York NY10017

© Ian W. Farmer
Softcover reprint of the hardcover 1st edition 1983
Photo typeset by Cotswold Typesetting Ltd, Gloucester
ISBN-13: 978-0-412-13980-2 e-ISBN-13: 978-94-009-5753-4
DOl: 10.1007/978-94-009-5753-4
This title is available in both hardbound and paperback editions. The
paperback edition is sold subject to the condition that it shall not, by
way of trade or otherwise, be lent, re-sold, hired out, or otherwise
circulated without the publisher's prior consent in any form of binding or
cover other than that in which it is published and without a similar
condition including this condition being imposed on the subsequent
purchaser.
All rights reserved. No part of this book may be reprinted, or
reproduced, or utilized in any form or by any electronic, mechanical or
other means, now known or hereafter invented, including photocopying
and recording, or in any information storage and retrieval system,
without permission in writing from the Publisher.

British Library Cataloguing in Publication Data

Farmer, I. W.
Engineering behaviour of rocks. - 2nd ed.
1. Rock mechanics
I. Title
624.1'5132 TA706

Library of Congress Cataloging in Publication Data

Farmer, I. W. (Ian William)


Engmeering behaviour of rocks.
Rev. ed. of: Engineering properties of rocks.
1st ed. 1968.
Bibliography: p.
Includes index.
1. Rock mechamcs. 2. Rocks-Testing. I. Title.
TA706.F36 1983 624.1'5132 82-19931
Contents

Preface vii

CHAPTER 1 ENGINEERING DESCRIPTION OF ROCKS


1.1 Rock testing 3
1.2 Uniaxial or unconfined strength 7
1.3 Empirical field and laboratory tests 14
1.4 Porosity and permeability 18
1.5 Discontinuous rock 24

CHAPTER 2 STRESS AND STRAIN


2.1 Stress at a point 33
2.2 Pore pressure and effective stress 37
2.3 Strain at a point 42
2.4 Representation of stress and strain 44
2.5 Relation between stress and strain 45
2.6 Geostatic stresses 51
2.7 Measurement of in situ stress 54

CHAPTER 3 ROCK DEFORMATION


3.1 Rock tests in compression 59
3.2 Rock deformation in compression 65
3.3 Mechanics of microfracture 69
3.4 Rock macrofracture 74
3.5 The complete rock deformation curve 77
vi Contents

CHAPTER 4 ROCK STRENGTH AND YIELD


4.1 Rock strength criteria 81
4.2 Yield criteria 85
4.3 The critical state concept 89
4.4 Triaxial testing 94
4.5 Axial and volumetric strain data 97
4.6 The Hvorslev surface in rocks 111

CHAPTER 5 TIME DEPENDENCY


5.1 Creep strain 120
5.2 Phenomenological models of creep 125
5.3 Time-dependent deformation 128
5.4 Time-dependent strength reduction 131
5.5 Cyclic loading 135
5.6 Rapid loading 139

CHAPTER 6 DISCONTINUITIES IN ROCK MASSES


6.1 Discontinuity measurement 145
6.2 Discontinuity orientation data 148
6.3 Shear resistance of a rock containing a
discontinuity 151
6.4 Shear resistance of a discontinuity 158
6.5 A critical state model for rock discontinuity
strength 165
6.6 Measurement of discontinuity shear resistance 167

CHAPTER 7 BEHA VIOUR OF ROCK MASSES


7.1 Discontinuity frequency 169
7.2 Rock mass classification systems 172
7.3 Rock mass strength criterion 184
7.4 The relevance of rock mass strength 189

REFERENCES 193
AUTHOR INDEX 201
SUBJECT INDEX 204
Preface

The first edition of this book was received more kindly than it
deserved by some, and with some scepticism by others. It set out to
present a simple, concise and reasonably comprehensive introduction
to some of the theoretical and empirical criteria which may be used to
define rock as a structural material.
The objectives - reinforced by the change in title - remain the
same, but the approach has been changed considerably and only one
or two sections have been retained from the first edition. The
particular aim in this edition is to provide a description of the
mechanical behaviour of rocks, based firmly upon experimental data,
which can be used to explain how rocks deform, fracture and yield,
and to show how this knowledge can be used in design. The major
emphasis is on the behaviour of rocks as materials, although in the
later chapters the behaviour of discontinuities in rocks, and the way in
which this can affect the behaviour of rock masses, is considered.
If this edition is an improvement on the first edition it reflects the
debt lowe to numerous people who have attempted to explain the
rudiments of the subject to me. I should like to thank Peter Attewell
and Roy Scott in particular. I should also like to thank Tony Price
and Mike Gilbert whose work at Newcastle I have used shamelessly.

I. W. Farmer
Newcastle upon Tyne, September 1982
1 Engineering
Description of
Rocks

Geologists recognize only one naturally occurring earth material


called rock. Engineers differentiate between rocks and soils, although
sometimes the dividing line is unclear. In particular, engineers
differentiate between the reactions of rocks and soils to the forces
imposed on them or in them by construction. The study of the
reaction of soils to these forces is called soil mechanics and the study of
the reaction of rocks is called rock mechanics.
Both rocks and soils are made up of mineral and organic particles.
In the former, the particles are generally bonded or cemented together
and an initial yield resistance must be overcome before they shear in
an unconfined state. Soils exhibit no real resistance to shear in an
unconfined state, and a very small energy input is required to
precipitate breakdown.
The test behaviour of soils can often be quite closely related to their
mineralogy. Although there are about 2000 minerals in the earth's
crust, silicates make up about 99 %of the total rock volume. The basic
silicate structure is such that only one form - quartz - can easily resist
the weathering processes to which rocks in the earth's environment
are subjected. Most of the others, originally constituents of igneous
rocks, are subject to chemical change during the weathering process
to form clay minerals, mainly hydrous aluminium silicates produced
by alteration of micas and feldspars.
Thus sediments - the end product of the weathering process - can
be viewed most simply as comprising two main materials, quartz and
clay mineral. Quartz particles tend to be blocky and equidimensional
and are generally larger than silt size (greater than 0.06 mm). A
sediment made up mainly of quartz particles would be described as

1
2 Engineering Behaviour of Rocks

silt, sand or gravel. Clay mineral particles tend to be small (less than
0.002 mm), flat and platy and a sediment containing a large
proportion of these would be referred to as a clay.
Sedimentary rocks have greater engineering interest because they
contain the weaker rocks. They result mainly from the compaction
and cementation of sediments, althou~h other processes including
recrystallization, replacement, differential solution and alteration
may occur. These processes are described by the general term
diagenesis and occur during changes in pressure and temperature as a
bed of sediments is buried beneath later sediments. The major
changes resulting from diagenesis may be summarized after
Krumbein (1942) as:
(a) Particle size - particularly in fine-grained sediments, may in-
crease through cementation, recrystallization or alteration.
(b) Particle shapes - may become more or less rounded through
solution or recrystallization.
(c) Particle orientations - may alter during compaction or re-
crystallization; usually this will be controlled by particle shape
and water content.
(d) Porosity and permeability - will usually be reduced through
compaction, cementation, solution and recrystallization.
(e) Structure - will be changed as the material changes from a free-
flowing sediment to a more brittle solid.
The mineralogical composition of a sediment may also change
during diagenesis but, it is generally true to say that the mineralogy of
sedimentary rocks is similar to that of the original sediments. Thus
shales and mudstones comprise predominantly clay mineral, and
sandstones and gritstones are usually a quartz aggregate cemented
together with a carbonate or clay mineral matrix. Depending on the
conditions of deposition, sandstones may of course have a high clay
mineral content and shales a high quartz content.
It is always important to define terminology correctly, and in the
case of soils and rocks this is confused by the different concepts of the
material and the mass. The material form - which is usually how the
soil or rock is delivered to the laboratory - may comprise either
discrete particles in the case of a soil, or an intact specimen held
together by interparticle bonding in the case of a rock. This is the
basic difference between rocks and soils as materials. A soil sample
may be held together by suction or other forces but it is inherently a
particulate system. The particles in a rock are cemented or bonded
Engineering Description of Rocks 3

together - in other words a rock has real rather than apparent


cohesion.
The massive form of rock or soil differs radically from the material
form. Soils are usually layered and their mechanical reactions and
ability to transmit water vary from layer to layer. Laboratory test
results can, however, usually be extrapolated, albeit with caution (see
for instance Rowe 1968) to the mass and it is not necessarily incorrect
to treat soil as a continuum. Rocks are often layered, but more
importantly they are fissured and jointed and this means that rock
masses may sometimes be controlled more in their reaction to forces
by the discrete nature ofthe fissured mass than by the properties ofthe
material.
Rock mechanics must therefore be defined as the study of rock
deformation and fracture in both its intact material form and as a
discontinuous mass. Nevertheless, through convention or otherwise,
rocks are usually described for engineering purposes through their
action as materials, and it is useful to start by considering some ofthe
simple tests to which rocks are subjected and which can be used to
define and compare their engineering reactions.

1.1 Rock testing

Since the title of the first edition of this book was Engineering
Properties of Rocks, it is important to start by qualifying the use ofthe
word 'properties'. The Shorter Oxford English Dictionary defines a
property as a 'characteristic quality of a person or thing'. It is
therefore correct to refer to the physical make-up or to the mechanical
reactions of rocks under test as properties. But it is also important to
remember that the properties of rocks obtained under laboratory test
conditions are related to the test conditions. For instance, a rock
cylinder tested in uniaxially unconfined compression will behave
differently ifits length/diameter ratio is 0.3 than ifits length/diameter
ratio is 3 (see Fig. 1.1).
There are therefore no fundamental mechanical properties of rocks
in the sense of material constants characteristic of a particular rock.
There are standard tests of various types which give useful indices of
rock properties for comparison with other rocks tested under similar
conditions. There are also some fundamental physical properties
which include mineralogical composition and phase relationships,
SIZE EFFECT
UD RAT IO 3
19mmDIA

l 00mm
DIA .
o ......,~~,.,....."'"-;,......,.--!

LlDRATIO 1
1z
:>:
'" ?O
'"
w
a:
:;;

LI D RAllO 1/3

o
123'567
&TRArN 1-/.1

SHAPE EF FECT
DIA 100mm
L10 "'] ·3

DIA.50 mm
LID 1/] . 3

DIA '9mm
LID 1/]_3

Figure 1.1 Influence of specimen shape expressed in terms of length/diameter


(L/D) ratio, and of specimen size expressed in terms of cylindrical specimen
diameter, on the deformation characteristics of Georgia Cherokee Marble
loaded in uniaxial compression at a constant strain rate of 10 - 5 S - 1 (after
Hudson et al. 1971).
Engineering Description of Rocks 5

and which although of limited engineering significance can be


considered material constants.
Decisions on test categories and specifications must of necessity be
arbitrary. They are also complicated by the existence of several
Qpecifications, the most complete being those ofthe American Society
for Testing and Materials (ASTM). Recently these have been
complemented by a series of 'Suggested Methods' (Brown 1981)
prepared by committees of the International Society for Rock
Mechanics (ISRM). The test categories recommended for standardiz-
ation are listed in Table 1.1.
N one of these tests will be considered in this book in greater detail
than is required for completeness, although some particular ap-
proaches to laboratory testing and field characterization will be
expanded later. In the case of laboratory tests, there are three main
aims in testing:

(a) To provide basic information on the physical properties and


mechanical reactions of the rock material.
(b) To classify or characterize the rock material by providing an
index which can be used to compare the particular rock with
other rocks.
(c) To provide information which can be used to design structures in
the rock.

Some of the tests described by Brown are too sophisticated for the
use to which they are put, and design in rock is often based more on
field measurement and empiricism than on laboratory test data.
Nowadays there is increasing emphasis on the need for large numbers
of quick in situ tests to give an indication of rock reactions rather than
for detailed information from a particular test.
Of the laboratory tests listed in Table 1.1, those dealing with
physical properties (l(a)(i) and (v)) will be referred to later. Of the
remaining tests, the most important are those describing strength and
deform ability. It must be stressed at once that all the laboratory tests
in Table 1.1 are index tests. None of them sets out, or is designed, to
provide information on the fundamental mechanical reactions or
deformation mechanisms of the material under load. Tests suitable
for this will be described in succeeding chapters. The present chapter
is concerned with standard tests and the index data from them, and in
particular:
(a) Uniaxial strength, ()cf - the greatest stress that a specimen can
6 Engineering Behaviour of Rocks

Table 1.1 Rock test categories (Brown 1981)

LABORATORY TESTS
(a) Classification
(i) Density; moisture content; porosity; absorption
(ii) Uniaxial tensile and compressive strength and deformation
characteristics
(iii) Anisotropy indices
(iv) Hardness; abrasiveness; attrition
(v) Permeability
(vi) Swelling and slake durability
(vii) Sonic velocity
(viii) Micro-petrographic descriptions
(b) Engineering design
(i) Triaxial compressive strength and deformation
characteristics
(ii) Direct shear tests
(iii) Time-dependent and plastic flow characteristics

2 FIELD OBSERVATIONS AND TESTS


(a) Characterization
(i) Discontinuity orientation; spacing; roughness; geometry;
etc.
(ii) Core recovery; RQD; fracture frequency
(iii) In situ sonic velocity
(iv) Geophysical borehole logging
(b) Engineering design
(i) Plate and borehole deform ability tests
(ii) Direct shear tests
(iii) Field permeability measurement
(iv) In situ rock stress determination
(v) Post-construction monitoring of rock movements
(vi) In situ uniaxial, biaxial and triaxial compressive strength

maintain when subjected to stress in a single direction, usually in


an axial direction in the case of a cylindrical specimen.
(b) Uniaxial deformation modulus, E - the ratio of normal stress to
strain for a material at a specified stress level when subjected to
stress in a single direction. Since the stress-strain curve is rarely
linear for earth materials, the standard value quoted is usually the
slope of a tangent to the curve at a stress equal to 0.50"cf.
(c) Poisson's ratio, v - the ratio between transverse and longitudinal
strain of a specimen subjected to uniaxial stress.
Engineering Description of Rocks 7

(d) Triaxial compressive strength, O"lf - the greatest compressive


stress that a specimen can maintain in the major principal stress
direction when subject to confining minor and/or intermediate
stresses.
(e) c and ¢ parameters - the cohesion and coefficient of internal
friction obtained from a series of triaxial tests at different
confining pressures and defined in Chapter 3.

1.2 Uniaxial or unconfined strength

Just as in concrete design the major criterion for specification is cube


strength, so in rock mechanics the most quoted index of mechanical
behaviour is unconfined strength. The major work on uniaxial
strength and testing is by Hawkes and Mellor (1970) and Hawkes et
al. (1973). It is particularly useful in that it points out the probable
effect on test results of irregularities which may occur during test, or
be present in the test specimen. The condition and size of the test
specimen is particularly important.
The ASTM Specification D2938 for unconfined compressive
strength tests requires that test specimens be right circular cylinders
with a diameter not less than NX (54 mm) core size and a
length/diameter ratio of 2.0-2.5. Where it is necessary to test smaller
cores, it is desirable that specimen diameters should be at least 10 times
the maximum mineral grain diameter. Hawkes and Mellor (1970)
suggest a figure of 20 times. The sides of the specimen should be
smooth and free from abrupt irregularities. The specimen ends should
be cut parallel to each other and normal to the longitudinal axis. The
ends should be ground and end lapped. Tolerances quoted in the
specification are:
(a) irregularities - all surfaces straight to 0.127 mm over the speci-
men length;
(b) end lapping - ends flat to 0.025 mm and perpendicular to the axis
to within 0.25°;
(c) height - five equally distributed measurements over the specimen
with a dial comparator should be within 0.051 mm.
Moisture content should normally represent field conditions.
The shape of the specimen is chosen as a cylinder rather than a
rectangular prism to avoid 'edge' effects. It is not, however, possible to
avoid end effects - particularly in uniaxial testing. Balla (1960)
8 Engineering Behaviour oj Rocks

considered the effects of total end restraint on cylindrical test


specimens with a length/diameter ratio of2. Using his results, Hawkes
and Mellor (1970) have shown that in specimens with a high ada 3
ratio, high deviatoric stress zones can be observed at the corners and
in the centre of the specimen. These are discussed in Chapter 3 (see
Fig. 3.2) in describing mechanisms of rock breakdown. In the present
context end effects are important insofar as they affect the results of
tests. A general guide to end effects is the empirical equation of Obert
et al. (1946):
a cf = a c1 (0.8 +0.2D/L) (1.1)
where a c1 is the measured uniaxial compressive strength of a
cylindrical specimen with a length/diameter (L/D) ratio of 1 and a cf is
the observed compressive strength.
The equation implies that within the experimental error for rock
testing in compression an L/D ratio greater than 2 is required for
ultimate values - and ideally a length between 2.5 and 3. Data
collected by Hawkes and Mellor from various sources are included in
Fig. 1.2.
Uniaxial tests can be carried out in various ways. Depending on the
direction of the applied force, strength may be measured in
N
e 300
Z
>:
I
t:;
z
..,
cr
~

~~
.....
V"

~ 200
...crVi
Vl

Q.

8
-'
,
«
x« \~
~~
:;: 100
...z
j

' "
' .....................
-- --=-----=-=----
L>J
cr ® ":::-:.:---
11.
Q.
<I
0
LID

Figure 1.2 Effect of length/diameter (L/D) ratio on the uniaxial compressive


strength of granite (1), dolomite (2), trachyte (3), sandstones and siltstones
(4-8) and saturated granite (9) from various sources (after Hawkes and
Mellor 1970).
Engineering Description of Rocks 9

compression, tension or shear. An arrangement for a laboratory direct


shear test is described by the Committee on Field Tests of the ISRM
(Brown 1981) and is based on the conventional laboratory shear box.
More accurate information can however be obtained from triaxial
tests (Chapter 4) and there is limited demand for shear testing of
intact rock. There is also a limited demand for tensile tests (see
Hawkes et at. 1973) which are rarely as repeatable as compression
tests. The exception is the point load test described in the next section.
The two types oftensile test normally carried out are a direct axial test
on a core possibly 'dumb-belled' to reduce end effects and fixed with a
resin adhesive into tubular end pieces which are freely mounted to
ensure a force direction parallel to the specimen axis; and indirect
tests where a short cylindrical specimen is loaded diametrally. This
induces high tensile stresses across the diameter between the loaded
points.
F or a standard, however, the ASTM uniaxial compressive strength
test is recommended. Prepared specimens are loaded to failure at a
rate less than 700 kN m - 2 S - 1 in a suitable testing machine. Then the
uniaxial compressive strength of the specimen is the maximum load
carried by the specimen during the test, divided by the cross-sectional
area, the result being expressed to the nearest 0.1 MN m - 2. The
report on the test should include the following:
(a) Lithological description of the rock.
(b) Source of sample including: geographical location, depth and
orientation, dates of sampling and testing and storage
environment.
(c) Specimen diameter and height.
(d) Moisture content and degree of saturation at time of test.
(e) Loading or deformation rate.
(f) Type of failure.
(g) Density, porosity and other available physical data.
Whilst not a material constant, the unconfined compressive
strength is a useful and widely quoted rock index property. Because
rocks vary so widely it is difficult to relate lithological or geological
descriptions to strength. Various classification schemes have been
suggested by Coates (1964) and Deere and Miller (1966), among
others, to differentiate strong from weak rocks. Ideally, such
classifications should be logarithmic, and a simple one suggested in
Attewell and Farmer (1976) on the basis of previous work is outlined
in Table 1.2.
10 EngIneering BehavIour of Rocb
Table 1.2 Classification of rocks on the basis of unconfined compressive
strength (after Attewell and Farmer 1976)

Strength Strength range


classification (MNm- 2 ) Typical rock types

Very weak 10-20 Weathered and weakly compacted


sedimentary rocks
Weak 20-40 Weakly cemented sedimentary rocks,
schists
Medium 40-80 Competent sedimentary rocks; some
strength low-density coarse igneous rocks
Strong 80-160 Competent igneous rocks; some
metamorphic rocks and fine-grained
sandstones
Very strong 160-320 Quartzites; dense, fine-grained
igneous rocks

The rock types suggested in Table 1.2 do not follow the


conventional type of geological classification, or conventional
geological terminology. The reasons are obvious - the factors which
affect the strength of rock specimens are not the same factors which
are used to construct geological classifications. Thus although
strength can be related to mineralogy - particularly quartz and clay
mineral content (see Fig. 1.3) - it is equally likely to be related to
N
I:
Z 400
:I:
CL AY MINERAL MATRICES
~ STRENGTH COk~E(T£O TO
CJ ElIMINATE EFFECTS 0
~ JOO CU "' ~R E S)ION
'"
V'
~
>
~ 700 ( ARBO ATE
i't.T ;Wf<;

g: ~
8
100

O ~ -
60--7~
O %
QUARTZ co Et.

Figure 1.3 Relation between quartz content and uniaxial compressive strength
for Coal Measures Sandstones (after Price 1966).
Engineering Description of Rocks 11

density and porosity, grain size and shape and anisotropy, and much
more likely to be affected by discontinuities present on a micro- or
macroscale.
The weakness in attempting to relate geological terminology to
mechanical indices can be seen by looking at some typical ranges of
uniaxial strengths for the common lithological groupings which are
reproduced in Table 1.3. They confirm the very large variation that
would be expected between gtrength and conven.ti~Ml gMlogical
descriptions. One of the few mechanical indices which does relate
quite closely to geological terminology is the ratio between deform-
ation modulus and compressive strength. Data collected by Hobbs
(1974), mainly for chalk and Trias rocks, are included in Fig. 1.4 and
data collected by Deere and Miller (1966) for typical igneous and
sedimentary rocks are included in Fig. 1.5.
The data on deformation modulus in Fig. 1.5 were computed from
field sonic velocity measurements. This is effectively the initial tangent
modulus at the confining pressure acting on the rock - usually zero if
the velocity is measured at near-surface exposures. In Fig. 1.4 the
deformation modulus is the secant modulus at customary bearing
pressure except where pressure or plate bearing test data are included.
Other deformation moduli commonly quoted are the secant and
tangent modulus at 50% of compressive strength. In a nonlinear
stress-axial strain plot (Fig. 1.6) there is a case for standardization,
but the variation is probably less than experimental error.
It can be seen from Figs 1.4 and 1.5 that on a plot of deformation
modulus against strength, lithologically similar materials tend to fall
into diagonally distributed groups in which large variations in
modulus and strength fall into a narrow ratio range. This allows for a
lithologically related index of brittleness.

Table 1.3 Typical rock strengths (MN m- 2 )

Compressive, (Jcr Tensile, (J Tf Shear, c

Granite 100-250 7-25 14-50


Dolerite 100-350 15-35 25-{i0
Basalt 100-300 10-30 20-{i0
Quartzite 150-300 10-30 20-{i0
Sandstone 20-170 4-25 8-40
Shale 5-100 2-10 3-30
Limestone 30-250 5-25 10-50
modulus rn tlO
1ft

o· 1 10 10~ 103 10'


COMPRESSIVE STR GTH Oq MN 1m2

UPPER (HALK
poroSltY<35% --r-----L_JA
10'
.
~ UPPER (HAL

~ MIDDLE (HALK ~
eo ~----
Q. LOWEll S" AL

z
>:

20 ' OC

Figure 1.4 Modulus ratio ranges for some chalk, Lias and Trias rocks,
together with some typical engineering materials and normally consolidated and
overconsolidated clays (after Hobbs 1974).
160 r------~-____.----____,_,

'-DI ABASE
80
M HIGH
.~ MODULU
RATIO GRA I ES
E 1.0 I----+---+--~-""'-"-+----;<-~

....
~ 20 1----+----.4-----r.(L--+--~
-'
=>
8
~ 10 1----+~-4-~----+----
25
;:::
«
E

~ 5
Cl

2-5
10 20 1.0 80 160 320
U IAXIAL COMPilESS I VE S RENGTH ac!,1N 1m 1

160

80
LIMESTONE &
OOLOI" ITE
1.0

~ 20
SA DSTON[

z 10
I:
SHALE

20 40 80 160 320

aCf 1m2

Figure 1.5 Modulus ratio ranges for typical igneous and sedimentary rocks
(qfter Deere and Miller 1966).
14 Engineering Behaviour of Rocks

MOOULUS
-INITIAL TANGE NT at 0
p . SECANT from 0 to P
q Q • TANGENT at q
STRESS R' CHORD f rom r to r'
p --- ---

STRAIN

Figure 1.6 Examples of deformation moduli from a nonlinear stress-strain


curve.

The reasons for the apparent brittleness represented by the high


modulus ratio of calcareous rocks and the less brittle behaviour of
shales and some sandstones is not easily explained. The ratio between
deformation modulus and strength over the complete stress-strain
curve to peak stress is essentially the inverse of the axial strain at
fracture, which will be related to the strength of bonding or adhesion
between particles or crystals in the rock.

t.3 Empirical field and laboratory tests

A simplification of uniaxial strength testing which can give a rapid


and accurate strength index in harder rocks is the point load index test.
It is becoming so widely used as a laboratory and field tool that it is
worth considering in some detail.
The ASTM standard test for compressive strength requires
expensive laboratory equipment and careful specimen preparation.
The point load index test (Franklin et al. 1971) requires a relatively
simple loading frame (Fig. 1.7) and can be carried out on virtually any
shape and size (see Brook 1977) of specimen. The test is usually
carried out on an unprepared core, obtained direct from drilling,
which is compressed between two conical points loaded from a simple
hydraulic hand pump. The core specimen should ideally be NX
(54 mm) size and have a length at least 1.5 times the diameter (D). The
core fails at a relatively low applied force (P) due to tensile failure over
the diametral area between the points - and in a similar way to the
indirect tensile test. The strength at failure is expressed as a point load
index 1s ' where
Engineering Description of Rocks 15

Figure 1.7 Point load strength index test apparatus.

(1.2)

A very close correlation has been shown by various workers (Fig. 1.8)
between Is and the uniaxial compressive strength O"cf:
(1.3)

where C is a constant equal to 24 for NX (54 mm diameter) cores.


Hoek and Bray (1974) suggest modified values of C = 17.5 for 20 mm
cores, 19 for 30 mm, 21 for 40 mm, 23 for 50 mm and 24.5 for 60 mm
cores. They also suggest that the test is only valid if a clean diametral
break occurs between cores. If the fracture runs to another plane as in
schistose rocks or if there are signs of cone penetration and crushing
as in weaker rocks, the results should be rejected. A simple field index
which can be obtained from surface exposures and which is
recommended by Deere and Miller (1966) is the Schmidt hammer
rebound number. The Schmidt hammer was originally developed to
determine the surface hardness of concrete. The mechanism of
operation is simple - a plunger, released by a spring, impacts against
the rock surface. The rebound distance ofthe plunger is read directly
from a numerical scale reading 0--80. Calibration curves have been
obtained experimentally relating the logarithm of the 28 day cube
16 Engineering Behaviour of Rocks

2 0 , - - - - - - . - - - - - . -- , - - - - , - -- - - - , - -- , - - - - - ,20
o 8ROCH &f A KlINI19121
x D'ANDREA ET AL [19651
+BIE IAWSKI [19741

S RO~
+ - - - - t - - - - - - - j l - NIORITE 15

+
.--;~"""'I1---+------i'0

o •

• 0

100 ISO -200 250 300 350


UNIAXIAL COMPRESSIVE STRE GTH aCf MN 1m'

Figure 1.8 Relation between point load strength index and uniaxial
compressive strength for N X core samples (after Bieniawski 1974).

strength of concrete to the rebound number in a linear manner and


these are generally accepted in the construction industry. A similar
chart was produced by Deere and Miller (1966) for rock (Fig. 1.9)and
the relationship has been confirmed by a number of workers (see
Rankilor 1974; Carter and Sneddon 1977).
The lack of confidence in the data results to a certain extent from
variability of test techniques and test conditions, and arises partly
from the function of the hammer and partly from the state of the rock
face. For instance, the hammer when worn can give a variable energy
impact. The hammer is designed to function in a horizontal direction
and corrections must be made for use at alternative angles. In weaker
rocks the hammer can give unreliable results and the rock surface can
affect results if it is wet, if it contains gouge or grit, or if a joint is near
the surface. A detailed statistical analysis of data by Poole and
Farmer (1980) indicated good correlation provided a minimum of
five impacts at each measurement point were carried out and the
highest rebound value at each point was selected.
Another test which is particularly useful in weaker rocks - and
those with a tendency to swell - is the slake durability test. It was
originally developed by Badger et al. (1956) to assess the potential
disintegration of Coal Measures Shales and seat earths during coal
preparation, and subsequently modified by Franklin et al. (1971) for
core logging. The slake durability index Id 2 describes the proportion
Engineering Description of Rocks 17
AVERAGE DISPERSION OF O CMPRESSIVE STRENGTli
FOR HOST ROC S - H 1m 2
o co co
~ ~ 2 ..,"'..;-."

: e...:z
400 ,"
350 11I1 IJ.Y 1- //
300 I LV' /. '//
250 ,- Vh V/- i
200 // ~ ~ {>
.. 150 IL /; ~ 8// V/; ~,"-
1/
~ - ~~~~ ~
E
:z:
1: -- - --c)- --
0100
-// /.
xc 90 - '/. r//
// V//. -/ '/ ' /
~ 80
~ 70 V/i/ V/,I,; '//: ~
tIi 60 /J W ~ ~/: ;¥
.... ;YA ~ ~ ~!
~ SO --I--
'"~ 40 A~~ [?" ~
"-

/~~,(/
1:

~ 30 !
<[

*
x
<[ I
~ 20
I
I
15 -
~V
I
I

10 I I
0 5 :0 ,15 fO ~S ?O ~5 ~O ~~ ~O 55 6,0

,
, 10 I I 20 I I 30 I ,40 I I ,50 ,60
10 20 30 40 50 60
2~1
0 ~~~joo-~-~~I--~~5~
IO--~~6b

~ ~ ~ ~
SCH MIDT HAMMER ITYPE l) RE~O U D NUM BER

Figure 1.9 Relation between Schmidt hammer rebound number and uniaxial
compressive strength of rocks (after Deere and Miller 1966).

of a 0.5 kg specimen of rock fragments remaining after rotation in a


sieve through a trough of water at 20 rev min - 1 for 20 minutes, with
an intermediate drying cycle. The residue in the sieve after the test
should comprise only non-clay mineral or very well compacted clay
mineral fragments. The quality of the rock is usually expressed in
terms of the index, thus:
18 Engineering Behaviour of Rocks

0-25% very low quality - mainly seatearths and shales


25-50% low quality
50-75% medium quality
75-95% high quality
95-100% very high quality - mainly fine grained siltstones and
sandstones
Although it should not be overemphasized Hassani and ScobIe (1981)
demonstrated a similar curvilinear relation between Id 2 and strength
to that illustrated in Fig. 1.3.

1.4 Porosity and permeability

Rocks do have some physical properties which can be used as an aid


to engineering description. The most useful are those which define the
phase relations in an element of the material. Rocks are usually
multiphase materials containing three phases (Fig. 1.10), a solid phase
(usually mineral particles), a liquid phase (usually water) and a gas
phase (usually air). The combined liquid and gas phases comprise the
voids in the material.
In a dry rock the liquid phase will be absent, and in a saturated
material the gas phase will be absent. Each phase has a weight (say w.,
Ww , Wg ) and each phase occupies a given volume (say v.,
Vw , Vg ).
Phase relationships can be defined through weight (moisture
content), volume (porosity, void ratio, saturation) or weight per unit
volume or unit weight. The term unit weight is preferred to the more
commonly used density since density is conventionally expressed in
mass units. Unit weight in common with other 'weights' in the present
book will be expressed in force units.
Typical phase relationships are:
Weight
moisture content
(1.4 )

Volume
porosity
Vv e
n=-- whence n=-- (1.5)
Vvv.+ l+e
Vg
WEIGHT [WI &
VV VOLUHEIVI SUBSCRIPTS :
SOLlD _~ ~~/,E:.. 1 I----=---l 1V", I 9 - GAS
----L w _ WATER
v - VOIDS
GAS -->' .;:;::::r~~~///~ V/////l
s - SOLIDS
Vs

FigUre 1.10 Representation of rock as a mu/tiphase material.


20 Engineering Behaviour of Rocks

void ratio
n
whence e=-- (1.6)
l+n
saturation (%)
(1.7)

where Vv = Vw + Vg
Unit weight
unit weight

(1.8)
dry unit weight
(1.9)

unit weight (solids)

Ys=-V
w. (1.10)
s
buoyant unit weight
(1.11)

Of these, the most commonly quoted in the case of rock are


porosity and dry unit weight. Typical values are given in Table 1.4 for
granular soils and rocks. In the case of clay soils where porosity is
dependent on moisture content, it is more usual to quote moisture
content and unit weight.
The data in Table 1.4 illustrate several useful points which although
simple have considerable importance in rock mechanics:
(a) The dry unit weight of a granular soil is related to the degree of
compaction of the soil. The maximum unit weight occurs at the
densest packing and in this state the compressibility of the soil
and its potential ability to transmit pore water will be at a
minimum.
(b) Sedimentary rocks - at a more advanced state of diagenesis than
soils - will have a much lower porosity, apart from some
cemented sandstones and low-density rocks such as chalk. In
igneous and metamorphic rocks, porosity is even lower. The
compressibility of rocks and their ability to transmit pore water
will be much lower than in soils, and in many rocks water flow
will be solely through fissures.
Engineering Description of Rocks 21

Table 1.4 Unit weight and porosity - rocks and soils

Dry unit
Porosity,n weight, Yd
(%) (kN m- 3 )

Granular soils* Ottawa Sand 44-33 14.5-17.3


(after Hough Uniform sand 50-29 13.0-18.5
1957) Uniform silt 52-29 12.6-18.5
Silty sand 47-23 13.7-20.0
Flne--coarse sand 49-17 IH-21.7
Micaceous sand 55-29 1l.9-18.9
Silt-sand-gravel 46-12 14.0-22.9
Sedimentary rocks Bunter Sandstones 17.1 2l.0
(after Rispin and Sandstone - Lazenby 13.4 22.6
Cooper 1972) Sandstone - Exeter 15.8 2l.9
Sandstone - Stain drop 14.2 22.3
Shale - Widdrington 3.7 28.0
Limestone - W olsingham 0.23 27.0
Chalk - Pits tone 24.6 18.8
Siliceous sandstone 3.2 24.3
Siliceous sandstone 5.6 23.4
Igneous and Quartzitic sandstone 0.71 26.6
metamorphic Quartzite 2.6 26.l
rocks Quartzite - Skye 0.98 26.3
(after Rispin and Granite - Creetown 0.75 26.5
Cooper 1972) Dolerite - Northumberland 0.44 28.9
Slate - Honister 0.77 27.4

"The porosIties and eqUIvalent Untt weIghts quoted for sods are III the loosest and
densest states. For both rocks and soils, porosIties and unit weIghts are for hand
samples.

(c) Because the tightness of packing will increase resistance to


breakdown - and indeed in most materials close contact will
create bonding and adhesion - it is possible to relate unit weight
to strength (see for instance D'Andrea et ai. 1965; Judd and Huber
1962).
The pore space in soils and in most rocks is interconnected, The
rate of discharge flow of a fluid through the pore passages is given by
Darcy's equation relating flow, Q, per unit discharge area, A, to
hydraulic gradient:
Q
-=V= k'I (1.12)
A
22 Engineering Behaviour of Rocks

where v is the discharge velocity, k is a constant known as the


coefficient of permeability of the fluid and i is the hydraulic gradient
expressed in dimensionless terms as the ratio between the fluid
driving head and the flow path length.
Darcy's equation is simply derived from Poiseuille's equation for
laminar fluid flow through a single tube, using the analogy of a soil or
rock matrix as a bundle of capillaries. Another form of the Darcy
equation is the Kozeny-Carman equation which relates the rate of
discharge flow to the hydraulic gradient, through the porosity, n, the
specific surface area of particles per unit volume of soil or rock, So,
and the fluid density (say Yw for water) and viscosity, 1'/, to give an
equation for k:

(1.13)

k has the same dimensions as discharge velocity and is constant for a


particular soil or rock (n, So) and fluid (Yw, 1'/). Values are usually
quoted for water flow at room temperature. Typical values in
Table 1.5 are for soil samples and intact rock samples tested in the
laboratory. Test methods for the former are described by Lambe
(1951) for the latter by Daw (1971).
The test data for rocks are incomplete but sufficient to illustrate a
basic premise. Under normal engineering hydraulic gradients a rock
or soil is considered impermeable if it has a coefficient of permeability
less than 10 - 7 m s - 1. Thus only very porous, coarse-grained rocks
like Bunter Sandstone can transmit water through their pore space in
sufficient quantities to cause engineering problems.

Table 1.5 Rock (data after Rispin and Cooper 1972) and soil
permeabilities

Soils I ntact rocks

Gravel >10- 2 ms- 1 Bunter Sandstone 2xlO- 7 ms- 1


Sands 10- 3_10- 5 m S-I Staindrop
4xlO- 8 ms- 1
Sandstone
Fine sands, 10- 5 -10- 7 m S-I
coarse silts Quartzite - Skye 2xlO- ll ms- 1
Slate - Honister 1.2 x 10- 13 m S-I
Silts 10- 7 -10- 9 m S-I
Granite - Creetown 10- 14 m S-I
Clays <1O- 9 ms- 1
Engineering Description of Rocks 23
However, rocks are discontinuous materials and most rocks have a
facility for water transmission along fissures or discontinuities. Hoek
and Bray (1974) quote typical permeabilities for jointed rock of
10 - 3 m s - 1, for open jointed rock of 10 - 2 m s - 1 and for heavily
fractured rock of 1 m s - 1. An expression for the permeability of
fissured or discontinuous rock analogous to equation (1.13) has been
quoted by Snow (1968) and Vaughan (1963) in the form:
k = NYw(53
(1.14)
1511
where N is the fracture frequency, (5 is the fracture width and N (5 is
equivalent to discontinuity porosity.
Fig. 1.11 relates fracture frequency and fissure width to per-
meability. Permeability can be measured directly in the field by some
form of borehole packer test, or by various falling or constant head
tests (see Hanna 1973).

0·1
N ,0 · oerm
N, 1 per '"
011--+--1 - ""'t-..:-''r"- N , 10 per m
N , 10
0 perm
·001
3 - ,:07-:--<'~'0-·-:-S-::'O.....·-=-6~10-::·?:--"O -8
1'' ' 0.-:-,--='='0-'·2:--:':-·=-

C()[ FFICIENT Of PERMEABI li TY km ls


Figure 1.11 Relation between fissure width, b, spacing, liN, and permeability,
k, measured in discontinuous rocks where water flow is principally through
fissures (after Attewell and Farmer 1976).

The usual procedure is to pump water at a standpipe pressure of


1 atmosphere (10 kN m - 2) into a, say 150 mm, washed-out borehole
between inflated hydraulic packers which may be 1-5 m apart
depending on fissure frequency. The test pressure should be reduced
in areas of weak or broken rock. The test duration should be at least
15 min, during which instantaneous and total flows are measured.
The approximate permeability can be computed in a radial direction
from the equation:
k=R.
LH (1.15)
24 Engineering Behaviour of Rocks

where Q is the average flow rate during the test, L is the borehole
length and H is the constant head of water.
More commonly and less accurately the permeability is expressed
in /ugeons, where 1 lugeon is equivalent to a test flow rate of 1 litre per
metre per minute at a pressure of 10 atmospheres; Ilugeon is
approximately 10 - 7 m s - 1, the limiting permeability of engineering
significance.

1.5 Discontinuous rock

The concept of fissure water flow in massive rocks is a useful


introduction to the importance of discontinuities in determining the
engineering reactions of rock masses. There is a tendency in rock
mechanics to concentrate too much on laboratory test data - with
some justification since field tests may be expensive and difficult to
control. There is nevertheless a large volume of work on the
behaviour of rock masses. This is considered in greater detail in
Chapter 6 and 7, but one or two concepts can usefully be introduced
here.
Terzaghi (1946) was one of the first to attempt a systematic
classification of massive rocks and to describe the possible mechan-
ical reactions of various rock types. A summary of the main rock
conditions is included in Table 1.6, as they might be found in tunnel
faces and sidewalls. In practice there will be no sharp boundaries
between the rock categories and the properties indicated by each of
the terms may vary between quite wide limits.
In order to make the classification useful for tunnels, Terzaghi
proposed roof and side loads which might be expected on tunnel
supports for each of the classification divisions.
These were based on Terzaghi's (1943) arching theory, which is
based on the assumption that a frictional material above a tunnel
roof, when allowed to move vertically downwards, mobilizes shear
resistance along vertical planes at the opening boundary (Fig. 1.12).
Then equating the disturbing forces (1'B dz - B dO" z) and resisting
forces (KoO"z tan 4» at the edges of an elemental strip of width dz, gives
(1.16a)
or
dO"z Ko
-dz = l' - 2-
B 0"
z
tan 4> (1.16b)
Table 1.6 Description of rock conditions (after Terzaghi 1946)

Hard and Contains neither joints nor hair cracks. If


intact rock fractured, it breaks across intact rock. After
blasting, spalls may drop ofT the roof for several
hours or days. At high stresses, spontaneous and
violent spalling of rock slabs from sides or roof
may occur
2 Hard Consists of individual strata with little or no
stratified or resistance against separation along the boun-
schistose rock daries between strata. The strata mayor may
not be weakened by transverse joints. In such
rock, spalling is quite common
3 Massive Contains joints and hair cracks, but the blocks
moderately between joints are intimately interlocked so that
jointed rock vertical walls do not require lateral support.
Spalling may occur
4 Moderately Consists of chemically intact or almost intact
and and very rock fragments which are entirely separated
5 blocky and from each other and imperfectly interlocked. In
seamy rock such rock, vertical wa1ls may require support
6 Crushed rock Comprises chemically intact rocI( having the
and and sand character of a crusher run and capable of
7 and gravel exerting considerable side pressure on tunnel
supports. If most or all of the fragments are as
small as fine sand grains and no recementation
has taken place, crushed rock below the water
table will exhibit the properties of a water-
bearing sand
8 Squeezing Slowly advances into the tunnel without percep-
and rock at tible volume increase. A prerequisite for squeeze
9 moderate and is a high percentage of microscopic and sub-
great depth microscopic particles of micaceous minerals or
of clay minerals with a low swelling capacity
10 Swelling Advances into the tunnel chiefly on account of
rock expansion. The capacity to swell seems to be
limited to those rocks which contain clay
minerals such as montmorillonite, with a high
swelling capacity
26 Engineering Behaviour of Rocks

dv.' _ B~dl

Figure 1.12 Representation of the stresses acting on an element, dz, of


frictional material above a two-dimensional opening when vertical movement is
allowed.

which solves for the boundary condition

to give

G"z = yB <p [1-ex P(-2Ko-=- tan <p)J+q exp(- 2K o-=- tan <p)
2Ko tan B B
(1.17)

which if zlB is large becomes


yB
G" = --'----- (1.18)
z 2Ko tan <p
where B is the arch width, equal to approximately S +H in the case of
tunnels with roughly equal span S and height H, y is the unit weight of
rock and Ko is the horizontal to vertical geostatic stress ratio (see
Chapter 2).
During tunnel construction it is impossible to avoid yielding of
arches, and through tests on arch legs to determine vertical pressures
exerted by the roof and sides, Terzaghi suggested maximum arch
pressures in terms of metres of rock arch of the magnitude given in
Table 1.7.
Similar classifications have been suggested by Stini (1950) and
Moye (1955) among others for tunnels, by Duncan and Goodman
(1968) for slopes and by Ward et al. (1968) for foundations. They are a
useful empirical guide which may be usefully developed from site
Engineering Description of Rocks 27
Table 1.7 Rock loads in terms of Terzaghi's (1946) classification

Classification I nitial rock Final rock


. number load (m rock) load (m rock)

1 o o
2 o 0.25S
3 o 0.5S
4 o 0.25S---O.35(S + H)
5 0-0.6(S + H)* (0.25-1.1)(S + H)
6 (0.5-1.2)(S + H) 1.1(S+H)
7 (1.0---1.2) (S + H) (1.1-1.4)(S+H)
8 (1.1-2.1)(S+H)
9 (2.1-4.5)(S + H)
10 up to 80

*S is tunnel span, H IS tunnel heIght

investigations as an aid for design and construction. They do


however, like Terzaghi's classification, say very little about the effect
of discontinuities on massive rock, except that the frictional material
(classifications 6 and 7) gives roughly the expected load. This is
reduced in the more intact rocks and increased in the squeezing rocks.
Of the methods which have been used to describe the effect of
discontinuities on the rock indices mentioned earlier, several may be
discussed at this stage. The first are size-strength relations, usually
wrongly attributed to Weibull (1952) who investigated the statistics
of tensile failure based on a study of crack extension. More correctly it
might be attributed to Evans and Pomeroy (1966) whose early work
on cubes of coal, a cleated and densely fissured material, showed a
strong relation between strength and size of the form:
(1.19)
where a is the length of the cube side and K is a constant.
Based on a 'weakest link' approach this is the basis for strength
computation of pillars in coal mines. It is also suggested (see
Bieniawski 1981) that this type of relation can be applied to other
rocks (Fig. 1.13). The confusion created by this is discussed in
Chapter 7 where it is shown that the strength-size relation is not a
feature of discontinuities in most rocks, but of the type of
deformation.
A more satisfactory basis for size effects, based on the number and
28 Engineering Behaviour of Rocks

1 nlr-~---.---.----r---~~
z 100\T
~ h..... IRON ORE
~ 5O~O( ( - - - " Jahns (19661

~ 20
Vi T DIOR ITE
0).. 0
0
~ A Pratt et at (19721

a 10 ,,\.o'-'.. ~o
!f
... 0 o- - - - - - -o:<i
;;t CO~L - · - ' - - -. -
~ Bienlowski 119671
z
:;;>
o OS 10 15 20 25
SIDE LENGT H. m

Figure 1.13 Relation between uniaxial compressive strength and side length of
a cube specimen for various rocks (after Bieniawski 1981).

size of discontinuities, are two descriptions of rock masses which will


be attributed here to Deere (1964) and Onodera (1963). These are
known as the rock quality designation or RQD, and the velocity index.
RQD like Terzaghi's classification was developed initially as a
method of assessing tunnel support requirements, although primarily
for rock bolting and shotcrete, rather than steel arch sets.
Although ostensibly a measure of discontinuity spacing, RQD was
originally conceived as a method of assessing the degree of weathering
of a recovered borehole core during logging, although it can also be
obtained from a scan-line. It expresses rock quality as the percentage
of the total length of a diamond drill core which occurs in intact
lengths greater than 4 inches (the original definition) or 0.1 m:

RQD= 100 'In x,L


1= 1
(1.20)

where x, is the length ofthe ith core length greater than 0.1 m, n is the
number of intact lengths greater than 0.1 m and L is the length of
borehole over which RQD is required. Typical values of RQD and
suggested quality bands are listed in Table 1.8.
RQD has been adopted as a standard for borehole logging by many
engineering geologists and forms the basis of some of the detailed
empirical methods of rock mass classification described in Chapter 7,
where its theoretical justification is also considered. The major
empirical importance of RQD however lies in some of its relations
with compressibility and seismic wave velocity in rock masses. These
Engineering Description of Rocks 29
Table 1.8 RQD quality bands

R~D Terzaghi
(%) Descnption classification

90-100 Excellent 1-3


75-90 Good 3-4
50-75 Fair 5
25-50 Poor 5-6
0-25 Very poor 6-7

are not altogether surprising since RQD, although it does not define
them exactly, is a measure of the openings in, and conditions of, the
discontinuity surfaces in the rock mass, and these will determine its
deformation and wave propagation characteristics.
Seismic wave velocities are discussed briefly in Chapter 5. It can
however be seen that, since the seismic compression wave velocity in
intact rocks has an average magnitude of 4500 m s - 1, in air a
magnitude of 350 m s - 1 and in water a magnitude of 1500 m s - \
pore or fissure space (even if saturated) will reduce massive rock wave
velocities significantly.
The relative effects of discontinuity or fracture porosity and pore
porosity are illustrated from work by Tourenq et al. (1971) in
Fig. 1.14. The quite wide variations of sonic velocity with discon-
tinuity porosity suggest a powerful tool for site investigation to
determine the extent of weathered or disturbed rock at an excavation
interface.

0·8

CI.
u

0·2 0·1. 06 08 1-0


£ MEAS/EcALC

Figure 1.14 The relative effects offissure porosity, nf (%), and pore porosity,
np (%), on the relation between ratios of measured and calculated P-wave
velocity, Cp , and measured and calculated deformation modulus, E (after
Tourenq et al. 1971).
30 Engineering Behaviour oj Rocks

1/
10

0
0,6

/
%
10
I
¥

V
0
[] ",">'
~~,,~
-:~
,,/~~<:s

~SSU~: RELATION

/ 20 40 60 100
ROCK QUALITY DESIGNATION ROD %
o MANHATTAN SCHIST
c RON IER MESA TUFF
'" HOCKENSOCK SILTSTONE

Figure 1.15 Relation between velocity index and RQD (after Deere et al.
1966).

The relation between velocity index defined as the square of the


ratio between the field and laboratory seismic velocities and RQD
(Fig. 1.15) was suggested by Onodera (1963) and confirmed by Deere
et al. (1966). Since the square of the seismic velocity is related to the
modulus of elasticity in elastic materials it would be expected that
RQD would also be related to the ratio of the field and laboratory
deformation moduli, and this can be shown to be the case. This ratio is
known as the rock massjactor,j, and is defined by Hobbs (1973, 1974)
as the ratio of the deform ability of the rock mass within any readily
identifiable lithological and structural component to that of the
deform ability of the intact rock comprising the component.
Comparative values of RQD, fracture frequency, velocity index andj
are given in Table 1.9.
The rock mass factor is particularly useful in estimating settlement
of foundations. Relations between RQD, fissure spacing and j are
illustrated in Fig. 1.16, replotted by Hobbs (1974) from work by
Deere (1966), Boughton (1968) and from various chalk sites.
Field deform ability measurements in all cases were based on large
plate tests.
Table 1.9 Relation between RQD and j

Fracture Velocity Mass


Quality RQD* frquency indext, factort,
classification* (%) (per metre) V~/V£t j

Very poor 0--25 > 15 0-0.2 0.2


Poor 25--50 15-8 0.2-0.4 0.2
Fair 50--75 8-5 0.4-0.6 0.2-0.5
Good 75-90 5-1 0.6-0.8 0.5-0.8
Excellent 90-100 1 0.8-1.0 0.8-1.0

*Deere and Miller (1966).


tCoon and Merritt (1970).
tVF IS the wave velocity III the field, VL the velocity III the laboratory.

10
2 1 1 0 DEERE
. . DWORSHAK DAM
08 I 2 N BOUGHTON TAS A IA
I FRESH FRACTURES

,
I 3. N BOUGHTO TASMANIA
I
0·6 ALTERATION PRODUCTS
o FRACTURES
j
04

o2
0
----===---
0 10 20 30 1.0
FRACTURES 1m

10 ...
KILLING HO~ME ,lINCS
08 2 ::=J MU
.
FORD . OR FOLK
NEWMARKET, CA BS .
0·6 4 -.--- UTTLEBROOK,K[ NT
(OP EN SHAFT I ~ ~~

0·1. ...."'~ ....\~~


k - - -..., 2 ,~<.. ~i'~
<0",<..° = 3

.
1).2

o 4 ···
o 10 30
FRAC TURE S , m

Figure 1.16 Relations between rock massfactor,j, and fracture frequency for
strong and weak rocks (after Hobbs 1974).
32 Engineering Behaviour of Rocks

Hobbs points out two main trends in the data:


(a) There is a near-exponential decay in j as fracture frequency
Increases.
(b) The scatter in the results is high in the case of weathered fractures
of variable openness; and lower in the case of clean tight fractures
in unweathered rock.
It can be seen from Fig. 1.16 that j is particularly sensitive to
variations offracture frequency in the range 2-10 fractures per metre'
(fracture spacings between 100 and 500 mm).
2 Stress
and Strain

The mechanical behaviour of rock is determined by its reaction,


characterized by deformation, to the force field of its physical
environment. This paraphrases a definition of rock mechanics quoted
by Judd (1964) as emanating from the US National Academy of
Sciences Committee on Rock Mechanics. Coates (1965), a civil
engineer who wrote one of the seminal books on the subject, put it
more directly as the 'study of the effects of forces on rocks'.
The forces acting on the rock induce in the rock body a state of
stress, a quantity with the dimension of force per unit area. The
deformations are usually expressed as strain, a dimensionless
quantity expressing deformation in terms of the original dimension.
The relations between stress and strain in an idealized material form
the basis of the mathematical theories of elasticity and plasticity
which are useful models of the behaviour of actual materials. Stress
and strain analysis is, of course, covered comprehensively in
numerous general texts and in several texts specific to rock
mechanics - particularly Jaeger (1969), Jaeger and Cook (1969) and
Obert and Duvall (1967). Some of the more important aspects, only,
are covered here.

2.1 Stress at a point

Stress at a point can be defined most simply by considering a plane of


random orientation in the body and of small area, (iA. Then if (iF
(Fig. 2.1), opposed by an equal and opposite force on the other side of
the elemental plane, is the resultant of all the forces exerted on (iA, the

33
34 Engineering Behaviour of Rocks

bA

Figure 2.1 Representation of stress at a point 0 in a plane of random


orientation.

stress at 0 in a direction OP normal to the plane can be defined as a


vector quantity:
bF
Pop = lim ~ (2.1)
,jA-->O uA
POP represents the stress vector in any direction in the body, and
because it is a vector quantity can be broken down into three
components - a directional component and a normal and tangential
stress component mutually at right angles or, ifthe orientation ofthe
plane is known, by three stress components. In order that stress can be
analysed easily, the plane on which it is acting may be related to a
system of rectangular coordinates. These are denoted as x, y, z in
engineering notation, where z is the vertical axis and x, yare mutually
perpendicular horizontal axes. If JA is placed in the yz plane (Fig. 2.2)

Figure 2.2 Representation of stress at a point 0 in the yz plane of a system of


Cartesian axes xyz.
Stress and Strain 35

in relation to such a system of Cartesian axes, then POP can be


represented by the components:
(2.2)

where ax, normal to c5A, is the (total) normal stress component and
acting in the yz plane are the tangential or shear stress
! xy' ! xz

components.
It should be noted that bF, POP and ax have all been given a positive
sign although they are shown as compressive in direction. This is
contrary to the convention adopted in most engineering stress
analysis, but is the convention commonly adopted in subsurface
engineering, and by most texts in soil and rock mechanics and
foundation engineering. Although in some design problems (par-
ticularly those involving underground concrete structures, such as
tunnel linings ) it can lead to inconsistencies, it can be justified because
the majority of underground stresses are normally compressive.
Another point that should be noted is that, unlike conventional
engineering materials which are generally impervious for engineering
design purposes, rocks must be considered pervious, porous and
saturated. The implications of this are considered in the following
section, but especially in problems involving rock masses the total
stress, a, may have to be modified to allow for porewater pressure.
If c5A is orientated so that it lies in the xz plane, the normal and
shear stress components at 0 will be:
(2.3)
and in the xy plane
(2.4 )
The nine components of stress which are therefore required fully to
define stress at a point:

(2.5)

may be represented in Fig. 2.3 as the forces acting on the faces of an


elemental cube. In fact, at a point in a body, constraint will impose
rotational equilibrium so that !xy = !yx, !xz=!zx and !yz =!zy and only
six stress components will be required to define fully stress at a point
in a body.
36 Engineering Behaviour of Rocks
(fz

cry

.-'-=>--+-- crx

zy

cr,

Figure 2.3 Representation of a three-dimensional stress field by stress


components on the faces of an elemental cube.

Stress is a tensor quantity. In defining stress it is necessary to specify


magnitude and two reference directions, made up of directional
components referred to a system of coordinates and using either six or
nine numbers. Force is a vector quantity made up of magnitude and
direction and requiring three numbers for definition. Scalar quan-
tities such as density or temperature require one number to specify
magnitude.
Although engineering stress notation will be used here, it should be
noted that it is common to use tensor-suffix notation, where stresses
are referred arbitrarily to directions 1,2, 3. Then the stress tensor may
be expressed as:
0'11 0'12

[
0'21 0'22 (2.6)

0'31 0'32

This can lead to some confusion with the engineering notation for
principal stresses. These are the normal stresses to the principal planes,
which are three orthogonal planes on each of which the stress is
uniquely normal and shear stresses are zero. The principal stresses are
designated 0' l' the major principal stress, 0'2' the intermediate principal
stress and 0'3' the minor principal stress. In the stress tensor of (2.6)
these become:
Stress and Strain 37

a1 0
[o az (2.7)
o 0
When the magnitudes and directions of the principal stresses are
known, it is convenient to use them as reference axes. It will be seen
that in rock mechanics it is convenient, although not necessarily
correct, to equate the major principal stress with the vertical or z
direction and to assume two equal horizontal stresses.

2.2 Pore pressure and effective stress

Rocks are not the homogeneous continuous materials idealized in


some theoretical approaches to rock mechanics. On a small scale they
are made up of particles or crystals cemented or bonded together with
pore space between. On a large scale a rock mass is replete with
discontinuities and fissures. On both a large and small scale rocks are
multiphase materials containing three phases, a solid phase (usually
mineral particles), a gas phase (usually air) and a liquid phase (usually
water). The relations between these phases (see Fig. 1.10) determine
the physical properties of rocks, such as unit weight, porosity and
moisture content. The relation between the pore pressure of air, u.,
and the pore pressure of water, uw ' in the gas or liquid phase and the
total stress, a, acting through the solid phase also determines the
effective stress, a', acting through the rock.
In the case of saturated soils where there is little bonding and
relatively low contact between discrete particles, Terzaghi (1925)
showed experimentally that the effective stress - by definition the
stress controlling compression or shear in rocks or soils - was given
by:
(2.8)
Bishop (1959) has also shown that, for the multiphase case in soils,
effective stress is given by:
(2.9)
where IX is a parameter related to the degree of saturation and equal to
unity in fully saturated soils.
In rocks, and in concrete, where contact between particles is greater
and the matrix is much less compressible and much less easily sheared
38 Engineering Behaviour of Rocks

than in soils, it might be thought that the concept of effective stress


should be modified. Skempton (1961a) showed that this was possible,
at the same time proposing a mechanism for the effective stress
concept through the relative compressibilities of the rock matrix and
the individual particles making up the matrix.
Skempton's analysis was based principally on his interpretation of
tests on quartzitic sandstone and Vermont Marble by Zismann (1933)
and Bridgeman (1928) reproduced in Fig. 2.4. These comprised two
types of tests on jacketed and unjacketed rock specimens. In the
former the rock was subjected to an all-round confining pressure, p,
through a membrane or jacket, in a drained condition so that the pore
pressure in the specimen was zero. Then for any small increase in
pressure from p to p + !:ip, the volume of the specimen will change
from V to V +!:i V and the compressibility, C, of the specimen will be
defined by:
- -!:iV = C up
A
= C up
A I
(2.10)
V
the total and effective stress being equal because the pore pressure, u,
is zero. As can be seen from Fig. 2.4, compressibility is not constant
over a large pressure range for a specimen of rock comprising more
than one phase, and will only reach a constant level when all phases
but the solid phase have been eliminated. This condition can be
simulated in an unjacketed test where the confining liquid is allowed
to penetrate the specimen so that each particle is compressed by a
hydrostatic pressure, u. Then for a similar increase in pressure, the
compressibility of the solid phase, Cs' will be defined by:
!:iV
-V = Cs !:iu=Cs !:!u w (2.11 )

where the hydrostatic pressure is equivalent to the porewater


pressure, Uw-
If it is assumed that interparticle friction is negligible, then the
combined volumetric strain of a saturated porous material will be
given by:
!:iV
- -V = )+ C
C(!:!p -!:!u w sw
!:!u (2.12)

which can be rearranged in the form:

- !:i; = c[ (1 - ~s) J
!:!p - !:!u w (2.13)
s
I
OUARTZITI( SANDST ONE
J
6 -,-- - - f-

~ I
JACKETED
~UARTI
L ~-
~r--== >-- j ~

~
'-
-~f
U JA(KETED - -

'"~
12 I
S
:z
T I
E
~
VEP. 0 T HARBLE
I-
:::; 0
in
Vi
~
'"0..1: \
\
8
0
w

\
6

4 I
~
CAL(ITE
QD

U JACKETED!
I
10 20 30 so 60

PRESSURE P ( JA CKETED I OR u (U JACKETED I

Figure 2.4 Compressibility tests carried out by Zismann (1933) and


Bridgeman (1928) on jacketed and unjacketed specimens of sandstone and
marble (after Skempton 1961 a).
40 Engineering Behaviour of Rocks

or

~p' = ~p - ( 1- ~s) ~uw (2.14 )

which is a restatement of Terzaghi's effective stress equation


(equation (2.8)).
Skempton also derived a similar equation based on the shear
strength of saturated materials in the form

a'=a-(l- atantan¢'If;)u W
(2.15)

where If; is the angle of intrinsic friction of the solid phase and ¢' the
angle of shearing resistance of the porous material.
The implications can be seen by examining data on compressibility
in Table 2.1. It can be seen that for soils the ratio Cs/C is negligible
and that equation (2.14) becomes a close approximation of equation
(2.8). For rocks the compressibility ratio is higher and it can be seen
that at low confining pressures, in particular, the effective stress
concept has reduced importance in intact rock. In fissured rock
masses however, where an analogy between discrete particles and
blocks bounded by discontinuities may be drawn, the concept has
considerable significance.
The concept of effective stress assumes a constant pore water

Table 2.1 Compressibilities at p=lO kN m- 2 (after Skempton 1961a);


Cw =480xlO- 6 m2 kN- 1

Compressibility
(m 2 kN - 1 X 106 )

Material C Cs Cs/C

Quartzitic Sandstone 58 27 0.46


Quincey Granite 75 19 0.25
Vermont Marble 175 14 0.08
Concrete 200 25 0.12
Dense sand 1800 27 0.0015
Loose sand 9000 27 0.000 3
London Clay 7500 20 0.00025
(overconsolidated)
Gosport Clay 60000 20 0.00003
Stress and Strain 41

pressure for all applied stresses. Before an element of rock or soil is


stressed, the porewater pressure can be estimated from its depth, ZM
below the phreatic surface:
(2.16)
However, in a saturated undrained rock or soil, changes in stress
can lead to changes in porewater pressure, which may significantly
affect shear resistance. Skempton (1954) using a similar line of
reasoning to that used in deriving equation (2.14) showed how pore
pressure increments induced by phase interaction during deform-
ation could be predicted.
In a hydrostatically stressed sample the relative distribution of
applied total stress increments between the rock and the pore water
will depend on the relative compressibilities of the rock matrix and the
pore water. Thus, for an applied total stress increment, l1a ( = l1a' + l1u
in an undrained state), stress-strain relations for the two phases may
be written:

1 I1V
l1a' = - -
C V

1 I1V
l1u=-- (2.17)
Cw nV
where C and Cw are respectively the matrix and water compres-
sibilities, V is the initial volume of rock and n is the porosity. Whence
(2.18)
and
l1u = l1a or (2.19)
1 + (nCw/C)

where
1
B=----
1 + (nCw/C)
is called the porewater pressure parameter B.
Values of B computed from the data in Table 2.1 are listed in
Table 2.2. They show that for compressible materials B = 1 and all
stress increments are transmitted as porewater pressure. In rocks
which have low compressibility, B can be as low as 0.25.
42 Engineering Behaviour of Rocks

Table 2.2 Porewater pressure parameters A and B in saturated rocks

A parametert
B parameter* (at failure)

Quartzite Sandstone 0.29 Rocks and heavily -0.5-0


Quincey Granite 0.34 overconsolidated clays
Vermont Marble 0.55 Lightly overconsolidated 0.3-0.7
Concrete 0.58 clays
Dense sand 0.99 Normally consolidated clays 0.7-1.3
Loose sand 1.00 Sensitive clays 1.5-2.5
London Clay 1.00 Loose fine sand 2-3
Gosport Clay 1.00

*Based on data from Table 2.1.


tBased on data from Lambe and Whitman (1979).

In an element of rock subjected to a deviator stress (0'1 -0'3) the


stresses acting on it can be considered to comprise a hydrostatic stress
(0' 3) and a deviator stress (0' 1 - 0' 3)' Then an increase in the applied
stresses equal to dO' 1, dO' 3 will give an increase in porewater pressures:
(2.20)
where the porewater pressure parameter A = 0.33 if the material
behaves elastically and Poisson's ratio v =0.33.
In practice A may be taken as 0.33 in rocks below failure levels. At
failure, however, in rocks and overconsolidated clays A is much lower
(Table 2.2) and will normally have a negative value - associated with
the dilation of the rock as it approaches failure - which will tend to
reduce porewater pressures unless there is rapid recharge. This can
lead to considerable anomalies in the testing and behaviour of
saturated rocks, particularly where rocks are in the presence of large
quantities of water at high pressure as limiting stresses are
approached. There would for instance be reason to expect very
different rock behaviour if submarine rocks were loaded at levels
approaching failure or fracture.

2.3 Strain at a point

Strain may be defined as the compression (positive) or extension


(negative) resulting from the application of external or body forces,
Stress and Strain 43
divided by the original dimension. In any discussion of strain it is
important to differentiate between infinitesimal and finite strain.
Infinitesimal strain can be defined as the strain resulting from
application of an increment of stress and can be treated mathe-
matically. Finite strain resulting from application of large stresses,
often over a prolonged time period and at high temperatures can only
sensibly be analysed in terms of the directions of stresses in the form of
the strain ellipse. This is the province ofthe structural geologist and is
covered extensively by Ramsay (1967).
The resultant effect of an incremental increase in the stress tensor is
to induce body displacement, rotation and shear. The latter can be
represented as two equal increments of shear strain in two
dimensions; their sum is equal to the angular distortion,), (Fig. 2.5) by
which shear strain is conventionally represented. The strain tensor
can therefore be represented by:

ex !Yxy ~')'xzl
[ ~')'yx ey 2')'yz (2.21)
2')' zx -!')' zy ez
and the principal strain tensor by

ro~l ~2 ~ 1
0 e3
(2.22)

The sum of the principal strains or principal strain increments is the


volumetric strain or dilation, a quantity of some significance in rock
mechanics:
(2.23)

Figure 2.5 Representation of infinitesimal shear strain. Note that angular


distortion y, leads to rigid body rotation !y. To represent in strain matrix, rotate
through ty to give equal increments of shear strain in x and y directions.
44 Engineering Behaviour of Rocks

2.4 Representation of stress and strain

Design problems in rock invariably involve assessments of resistance


to shearing along a plane normal to a three-dimensional stress field. It
is useful therefore to consider two of the most common methods of
representing graphically the stress or strain tensor.
Consider in Fig. 2.6 the stresses acting on a plane of area, a, normal
e
to the xz plane whose normal is inclined at angle to the z axis, where
az is assumed greater than ax. Then the forces acting on any plane
passing through 0 will be given by:
ana=(a z cos 8-T zx sin 8)a cos 8+(ax sin 8-Txz cos 8)a sin 8
(2.24 )
ra = (a z sin 8 + Tzx cos 8)a cos e- (a x cos 8 + Txz sin 8)a sin 8
(2.25)

<Jx (J'x
x _._._. - _·_· - x

(J'z

z
COMBINED STRESS IN A PLANE

Figure 2.6 Representation ofa plane whose normal is inclined at an angle () to


the z-direction in a two-dimensional stress field.
Stress and Strain 45

which since 'zx='xz become:


O"n = O"z cos 2 8 + 0"x sin 2 8 - 2, xz sin 8 cos 8 (2.26)

,= (O"z - O"x)sin 8 cos 8-'xAsin 2 8-cos 2 8) (2.27)


Then if the principal stress axes are substituted for the x, Z axes so that
0"z = 0" l' 0" x = 0" 3 and, xz =, zx = 0, equations (2.26) and (2.27) become:

O"n = 0" 1 cos 2 8 + 0" 3 sin 2 8 = -!-(O" 1 + 0" 3)+ -!-(O" 1 - 0" 3)COS 28 (2.28)

(2.29)
This series of planes and the stresses acting on them can be
represented in " 0" nspace as a circle of radius -!-( 0" 1- 0" 3) with its centre
at -!-(O"l +0"3)' The circumference of the circle represents the normal
and shear stresses on all the planes normal to the plane containing 0" 1
and 0"3 and having a normal inclined at an angle 8 to the 0" 1 direction.
This particular representation of stress (Fig. 2.7) in two dimensions,
known as the Mohr circle, is rather clumsy if a series of different
stresses are required to plot graphically a stress path in a rock body.

--- -- -- - - -----~_T'""~

~ • 0"1-11J Co. 1 9
1 2

Figure 2.7 Mohr circle representation of stress on a plane.


46 Engineering Behaviour of Rocks

This can be done more conveniently by single point representation by


the stress invariants in two dimensions: the circular stress
P=!(O"I +0"3) and the deviator stress q=t(O"I -0"3)' These are also,
conveniently the 'top' points on the circle construction, and represent
the maximum stresses on a plane with its normal inclined at 45° to the
major principal stress. A typical stress path for a triaxial test in
compression is illustrated in Fig. 2.8.
Plane strain at a point can be represented in the same way as stress.
Thus the normal and shear strains on a plane normal to the direction
of two principal strains, the normal of which is inclined at an angle ()
to the direction of major principal strain, will be given by:
(2.30)

zY-z
1_1([;1 -[;2 )'2()
sm (2.31 )
which may be represented on a strain circle in exactly the same way as
stress in Fig. 2.7.
The notation is positive for compression and negative for
expansion.

2.5 Relation between stress and strain

The Cartesian tensor for describing stress at a point and comprising


three principal components (equation (2.7)) can be divided into one

J J
spherical (p) and three deviatoric (q) tensors:

r" " ~{ 1 J+q{ 1

+q, r~l 011+q, r1~l J (2.32)


where
P=1(0"1 +0"2+0"3)
ql =1 (0"2 -0"3)
q2 =1 (0"3 -0"1)
Q3=1(0"1-O"Z)
Stress and Strain 47
A similar notation can be used to represent the resultant strain
tensor by substituting principal, spherical and deviatoric strains. If
stresses and strains are assumed to occur in a linear elastic and
isotropic material, then the principal axes of stress and strain must
coincide, and in that case the properties of the material can be
described by two fundamental material properties, the bulk modulus,
K, and shear modulus, G, where:

(2.33)

and shear or deviatoric strain


q
Y=2G (2.34 )

In practice, engineering design is based on elastic 'properties'


obtained mainly from uniaxial compression testing (see
Chapter 1) - the modulus of elasticity, E, and Poisson's ratio, v, where:

(2.35)

and

(2.36)

which can be written in tensor notation as:

and combined with the general stress and strain tensors and
equations (2.33) and (2.34) to give:
1 1 1
(2.38)
E= 9K + 3G
v
(2.39)
E 6G 9K
q'
A B
p,-ai.ax'
- -r- 3cr~
-2-
Ko STRESS PATH
q' = rJ l-
2
0'; : rj'
'2
WITH I CREAS I (j
OEPT -
ax: Ko cr
z
Ko : 0, 5

~ _ _L -_ _L -_ _ _ _ _ _ ~ _ __ _ _ _ _ _ _ _ _ _ _ _ ~ ____ ~ p'

q'
TR I AXIAL TES T S !lESS
PA~ -err CON STANT ,
ff,' INCREASING

L -______~__~__~___ L _ __ L _~_ _ _~ _ _ ~ p '

Figure 2.8 Stress paths representing increasing stress with depth (Ko stress
path) and increasing axial or vertical stress (O"~) during a triaxial test subject to
an initial balanced compression stress (0";) which is subsequently maintained as
a radial confining pressure (after Attewell and Farmer 1976),
Stress and Strain 49

Schofield and Wroth (1968) point out that this demonstrates that
in an elastic material, axial compression of l/E is only partly caused
by spherical compression (1/9K) and is mainly caused by distortion
due to shearing (1/3G). Lateral deformation (v/E) is the difference
between shearing distortion (1/6G) and spherical compression
(1/9K). Their conclusion - important to stress analysis in rocks and
soils - is that an isotropic elastic body is not capable of reduction to a
set of three orthogonal coil springs. This leads to their subsequent
choice of equations describing spherical compression and shear as the
basis for their critical state concept of soil mechanics. The adaptation
of such an approach to rock mechanics based on dilation and shear
would be relatively simple, and is outlined in Chapter 4. It has the
obvious advantage of providing an analytical approach to rock
behaviour based on rock properties which correctly describe both the
idealized behaviour and the actual mechanism of deformation of
rocks under stress. It has the possibly major disadvantage that it does
not describe deformation in conventional engineering terms.
In conventional engineering terms, deformation is usually reduced
to three orthogonal springs. Then strain at a point can be related to
the deformation and induced lateral deformation produced in each
principal stress direction if the material is assumed to be isotropic and
linearly elastic. Thus it is possible to obtain a conventional statement
of equation (2.37) extended to the other two directions:
a1 va 2 va 3
I> ------
1- E E E
(2.40)

a2 va 1 va 3
I> ------
2- E E E
(2.41 )

a 3 va 1 va 2
I> ------
3- E E E'
(2.42)

The volumetric strain is the sum of the directional strains and is


given by:

where v, Poisson's ratio, is the ratio between strain induced in a


direction perpendicular to the applied stress and the strain in the
direction of applied stress.
50 Engineering Behaviour of Rocks

It is useful at this stage to consider values of Poisson's ratio, which


can sometimes be misleading when measured, and which illustrate the
weakness of the conventional approach to rock deformation. Take for
instance a conventional description of deformation for a uniaxially
compressed elastic material (Fig. 2.9) in which an idealized lattice
structure of spherical atoms is bonded together in rhombically
arranged rows. Then if the structure is contracted in one direction

Uz


+
I

z [x:o-2&x
x

[:~~f
[z= _2&z
z

tOz

tOz
I

t--- - X

where ox :1
thus 5z ./j
~x : 1'3="\1
Ey

Figure 2.9 Idealized representation of elastic deformation of a material made


up from a lattice structure of spherical atoms, and subjected to a uniaxial stress,
(1z·
Stress and Strain 51

with a resultant expansion in the opposite direction, it can be shown


that v=0.33, and by substituting in equation (2.43),
~V 1 (ITt +IT 2 +IT 3)
- ~

(2.44 )
V E 3

or, in other words the condition where the bulk modulus, K, is equal
to the modulus of elasticity, E, and there is the closest approach to
pure compression.
But it can be shown (see Chapter 4) that rocks approach a plastic or
critical state at such high confining and deviatoric stress levels, and in
this case ~ VIV = 0 and by substitution in equation (2.43), v = 0.5.
Before this level there may be dilation and in that case v> 0.5.
However, in many tests, particularly short-term uniaxial tests, at low
stress levels, rocks exhibit a high degree of lateral stiffness and in this
case v<0.2. This is a common result from testing and may be due
principally to recompression of pore or fissure space existing
originally or formed by stress release during coring. It may result from
a time element inherent in rock deformation processes and related to
the discrete nature of the rock structure. The application of data
based on such testing to design in rocks at high stress levels can be
difficult to justify, and can lead to significant error. This can be
illustrated by considering the case of the computation and measure-
ment of geostatic stresses.

2.6 Geostatic stresses

The usual assumption made in engineering is that the principal


stresses below the earth's surface are in the vertical and horizontal
directions and that the vertical (total) stress (IT z ) at a point is equal to
the weight of a vertical column of rock above that point:

(ITv=IT t = )ITz = J: ydz=yz (2.45)

where y is the rock unit weight and z is the depth below the surface.
Brown and Hoek (1978) have collected data on the relation
between vertical and horizontal stress and depth from various
countries, obtained using a wide range of measurement and
computational techniques. The relation between vertical stress and
depth is reproduced in Fig. 2.10 and confirms the form of equation
VERTICAL STRESS (h i'1N 1m2
10 20 30 40 50 60 70

500

" "
00 _

1000
-"
- 21
-- .-
e 1500 • 103

:r;;
"72
~
Cl
2000 LEGENO 63 "

• AUSTRALI A
... CANAOA
• 108
2500 I " U S.A .
• SOUTHERN AfRICA

o S(ANOINAVI A " 79
3000 f- .. OTHER REGIONS

Figure 2.10 Relation between vertical geostatic (total) stress and depth from data collectedfrom various sources (after Br~wn and
Hoek 1978). The numbers which refer to spec(jic case histories listed in the reference are not identified here, but are retained for
comparison with Fig. 2.11.
Stress and Strain 53

(2.45), the average magnitude of y being approximately 27 kN m - 3.


There is considerable scatter at shallow depths, this being attributed
to the relative insensitivity of the measuring techniques. There can
also of course be unusually high vertical stresses in the vicinity of
geological or topographical structures.
Equation (2.45) refers to the total stresses. In saturated porous or
fissured rock - the rule rather than the exception at depths up to
500 m - the relation should be written in terms of effective stress:
(2.46)
where Ys is the submerged unit weight of rock equal to y - Yw, Yw is the
unit weight of water and hw is the depth of water below the phreatic
surface.
A similar convention should always be used in attempting to relate
horizontal and vertical stresses because the hydrostatic pressure will
be equal in all directions, and the phreatic and ground surfaces will
not necessarily coincide. The conventional assumption is that the
horizontal stresses are equal in the whole horizontal plane and are
related to the vertical stress in the form:
(2.47)
where Ko is always expressed in terms of effective stresses and is
variously known as the coefficient of earth pressure at rest or the
coefficient of geostatic stress.
Hit is assumed that, at depth, rocks subjected to high cover stresses
are subjected in geological time to some form of stress relaxation, then
it is a valid assumption that Ko = 1, and the stresses can be described
as hydrostatic or more correctly as lithostatic and O'~ = O'~ = 0';. Jaeger
and Cook (1969) quote Talobre (1957) as describing this statement as
Heim's rule. It is equally true in the case of total stresses in saturated
rocks or soils, where the piezometric and ground surfaces coincide.
Hit is assumed that the rock behaves elastically, and that the earth's
crust offers just sufficient restraint to prevent horizontal deformation,
then substituting in equations (2.41) and (2.42) for £2=£3=0,
0'2=0'3=O'x and 0'1 =O'z, gives:
v
O'x = - - O'z (2.48 )
I-v
giving in the effective stress case:
v
Ko=--
I-v
54 Engineering Behaviour of Rocks

a relation originally proposed by Terzaghi and Richart (1952). Since


it can be shown (Fig. 2.9) that in the elastic case v = 0.33, then if rocks
behave elastically Ko should be equal to 0.5, and this is a common
assumption. In practice Heim's rule that Ko = 1 may be nearer the
truth and at shallow depths it appears that rocks may have much in
common with overconsolidated soils where Skempton (1961 b) has
shown that very high Ko values, decreasing with depth, occur in
shallow deposits. The high horizontal stresses are residual stresses
remaining after the erosion ofthick overburden and can be estimated
in soils from the overconsolidation ratio determined from uniaxially
confined compression tests. There is also evidence for residual stresses
in shallow rocks. Fig. 2.11 (from Brown and Hoek 1978) collects
data from the same sources as Fig. 2.10 in the form of Ko against
depth. It follows the general trend of similar data collected by
Jamison and Cook (1979), which is reproduced in Fig. 2.12 and which
illustrates the significant effect that thrust fault structures have on the
horizontal geostatic stress.
Although it is possible to fit limit curves, it is better to comment
generally on the data. This shows that at shallow depths, very high
horizontal stress can occur, particularly in rocks affected by thrust
faulting. For depths greater than 1000 m the horizontal and vertical
stresses tend to be similar. This is not surprising, for some relaxation
of high deviatoric stresses would be expected in most rocks over a
period of geological time. It must however be stressed that many of
the data are based on average horizontal stresses and many are of
doubtful reliability. The major conclusion might be that the data
indicate how difficult it is to predict in situ stress distributions
accurately by theoretical or empirical methods and how important it
is either to measure in situ stresses or to assess carefully structural
conditions affecting in situ stress as part of any design programme.
Another point is that where stress imbalances may be expected to be
high - in relatively shallow rocks - the strength of the rock may be
sufficient to reduce their importance, particularly compared with the
effect of discontinuities. At greater depths the likelihood of the
existence of high Ko values is reduced.

2.7 Measurement of in situ stress

The principles of in situ stress measurement in rock and the basic


mathematics for computation of stress based on elastic stress and
K = O'~ a.
o a;-
o O.S 1.0 1.5 2.0 2.5 3.0
I i i i i :L

- .... - ~~ -- ~ -- . - ..... ~
....... j; ,j-. . " ,.;. • ~ I •
/'.
/;. ••• "'-.... 1
-.. • .* 0 ....
. • o
soo I • ••• 0
I . "......
I ••• o · o • ..- ... -
I o ••" • .... .....'-
I 0 •• • 0 0 ....-
1000 I . • /....-
I • '" 77 •• . 27 • /'. ..... · 22
I 102 .
. /'
I · 104 ;21
e 1500 I . 105 ~6
I . • ,03
/
74'" . 12 /... K - 1500 • 0. 5
I 73 ,. o z
i!' I TIl *114 I
~
o
I "'63 I
2000 /
I
I
I
• • '06 I
I
I 112 . 106
2500
107
I . 'I
I
I K • lQQ.
• z • 03
"'79
Figure 2.11 Relation between horizontal geostatic (total) stress and depth from data collected from various sources (after Bro..wn and
Hoek 1978); see Fig. 2.1Ofor legend.
cr '
O~
K• = ~
o 5 6 B 9 10 II 12 13 14
1.-4 291t.~3' X27 I1 X 24 " 21 23"
100 0" "32
1.30 2 36 X X'?
,,'5 '8,/37 "28
12~8, - - --
x38 "19
39 ~~, 25" x31. "20
--- ---
o 60~
L6 Ot.OOlO x 33
4'° 01 3
,,-
500 r ...
... K. • - liQQ. • 05
/ 1
I
I
01.1 x35 I
I
01.8 I
400 0\5 I
1.50 °50 I
1000 ~ 26 " I
I
I
e I
I
:r: I
>-
"'-
..... I
0 I
I

1500 0 NORMAL FAUL TlNG CONDITIONS


x TH RUS T FAULTI NG CONDITIONS
0 STRIKE·SLlP FAULTING CONDITIONS

0 42
I
moo
Figure 2.12 Effect of geological structures on Ko values (after Jamison and Cook 1979). The numbers refer to case histories identified
in the reference.
Stress and Strain 57
strain analysis are summarized by Leeman (1964), Jaeger and Cook
(1969) and Roberts (1977), and are listed in Table 2.3. Leeman also
describes in detail some of the devices which have been developed,
particularly for measuring borehole deformation.
The obvious drawbacks to the methods of stress determination are:
(a) The mathematical treatments are based on assumptions of
elasticity, homogeneity and isotropy, which are not always viable.
Adaptations of the equations to allow for these are not always
satisfactory.
(b) A value of Poisson's ratio must be assumed in many of the
constitutive equations and this strongly influences the stress
values computed. Conditions in situ are so different to those in the
laboratory that v cannot be measured satisfactorily.
(c) All the methods require a borehole or a slot to be formed which is
itself instrumental in destressing the ground.
The only methods which appear conceptually sound in Table 2.3
are the inclusion stress-meter which will only measure changes in
stress, and hydrofracture which has the disadvantage of inaccurate
determination of fracture onset and a requirement to estimate tensile
strength. Against this, some of the more reliable data on in situ
measurement appear to have been obtained in this way.
It is, nevertheless, vital that in the design of complex underground
excavations, there is information on in situ stresses. The importance of
persevering with and improving existing methods should be
emphasized.

Table 2.3 Methods of in situ stress measurement

Method Description Reference

Flat-jack Measurement of stresses at exca- Jaeger and·


vation surface. Cut slot measuring Cook (1969)
convergence. Restore convergence
with hydraulic jack, whose press-
ure measures vertical stress re-
lieved by cutting slot
58 Engineering Behaviour of Rocks

Table 2.3 (continued)

Method Description Reference

Borehole Measure changes, b, in diameter, a, Leeman (1964)


deformation of borehole caused by overcoring.
Then at angle 0 to major principal
stress az> in borehole parallel to
intermediate principal stress a y:
2a
b = If [a z + ax - vay + 2(az - ax)
(1 - v2 )cos 20]

Doorstopper Measure strains induced in e.r.s. Leeman (1964)


gauge rosette cemented to end of
borehole and overcored. Then if
the borehole is in the y-direction:
1
tX=}i [a(a x -vaz )-b(l-v)ay]
1
tz =}i [a(a z -vax )-b(l-v)ay]

where a, b are constants


Borehole Measure strains induced in three Leeman and
strain e.r.s. gauge rosettes cemented Hayes (1966)
radially around a borehole and
overcored. Mathematics for calcu-
lation of stresses, based on elastic
analysis, is given in reference
Inclusion Device with higher modulus than Roberts (1977)
stress-meter rock cemented directly into a bo-
rehole. Then change in stress in
rock in a direction normal to the
borehole is given directly by:
I1a 2k+ 1 2
--=~~-+- as k-+oo
~ao 3k 3
where ~ao is the change in stress of
the inclusion and k is the ratio of
inclusion to rock modulus
Hydrofracture Measure borehole pressure Po at Haimson
which hydro fractures are initiated (1978)
in a direction parallel to the minor
principal stress. Directions can be
determined by overcoring. At
hydrofracture initiation:
Po;' aT[ + (3ay - a x)
3 Rock
Deformation

In the case of rocks, terms used in conventional stress analysis to


define stress-strain relations in ideal materials - such as elasticity,
plasticity and viscoelasticity - are not easily justified. Fairhurst, in his
foreword to Jaeger and Cook (1969), summarizes the position very
well:
... problems of rock behaviour involve considerably more than
selecting appropriate elastic constants and strength parameters
and inserting them into elementary theories of continuum
mechanics ... Although rock obeys the same laws of mechanics as
other materials, there are sufficient differences in behaviour, and
emphasis in methods of approach, to warrant the distinctive term
'rock mechanics' ... Inelastic deformations are, in fact, often of
major significance in determining the stability and safety of
structures in rock.
It is important, therefore, in order to understand the behaviour of
rock, to examine in detail what happens during some of the simple
tests carried out on rock, and to try to explain the mechanism of
deformation.

3.1. Rock tests in compression

The philosophy of rock testing has been outlined in Chapter 1, and


the detailed methods are described in the relevant American Society
for Testing and Materials Specifications and British Standards and in
the Suggested Methods of the International Society of Rock

59
60 Engineering Behaviour of Rocks

Mechanics (Brown 1981). Whilst it is not necessary to describe test


methods in detail, some information is necessary in order to explain
fully the mechanism of rock deformation. This includes, particularly,
the choice of test specimen geometry and the choice of testing machine.
The choice of a cylindrical specimen to minimize edge effects has been
mentioned in Chapter 1. It is not, however, possible to avoid end
effects ~ even where 'frictionless' platens are used. Balla (1960)
considered the effects of end restraint on cylindrical test specimens.
He characterized the stress condition of the test specimen in terms of
the tangential stress (equal to (2/3)1/2 times the octahedral shear
stress) for the uniaxial and triaxial compressive stress regimes
illustrated in Fig. 3.1. In the specimens with a high (J d(J 3 ratio, high
deviatoric stress zones can be observed at the corners and in the
centre of a specimen. In the extreme case of uniaxial compression,
Hawkes and Mellor (1970) have computed the stress distribution in
terms ofthe McLintock~Walsh failure criterion (equation (3.15)) and
the contours in Fig. 3.2 indicate zones of high tensile stress at the
centre and corners of the specimen. The effect of end restraint on
specimen behaviour is discussed later in this chapter. For the present,
it is sufficient to note the potential for any description of rock
deformation based on extension of fractures in a tensile stress field.
According to ASTM Specification 02938 the testing machine for
compression testing should have sufficient capacity to apply a load
continuously and evenly and at a constant rate so that failure occurs
between 5 and 15 min after commencement of loading. This should
give results which are reasonably free from the long-term and short-
term loading effects discussed in Chapter 5.
Loading platens should have a Rockwell hardness greater than
HRC 58 and one should be spherically seated, the other rigid. The
spherical seating should be between one and two times the diamter of
the specimen. The present book is not the place to discuss in detail the
design of testing machines, but one point of considerable importance
must be made. This concerns the importance of using a machine with
a loading and unloading characteristic stiffer than the specimen under
test.
Most testing machines of the conventional type use force as the
independent variable. Collapse of the specimen is achieved by
increasing the force beyond the maximum load-bearing capacity of
the specimen ~ with catastrophic results as a large amount of stored
energy is released into the specimen. If a machine is designed so that
displacement is the independent variable and force the dependent
~~
! '·0
" 0 ·30
0·] 0
o8 0.8 0.' 0

06 0·6

o I. 0·1.

02 0 ·]

o L-.\-.,....-..,-..I,.--l 0
o 0·2 0.. O~ 0-8 '·0 o 0·] 0-1. 0·6 0·8 1-0 0 0·2 0·1. 0·6 0·8 .0
r/R r/R r/R

10",0 6 .0 0""10 1

~ '·0 050

0·6

0 ·1.

0 ·] 0·2

o L--r-'-~~~---.J O L-~~...J,.... __
o 0·] 0·1. 0·6 o·e 1·0 o 0·2 0 1· . 0·6 o·a '·0 o 0 ·] 0·1. 0·6 Oil l-O
r/ R r/R r/R

Figure 3.1 Tangential stress (equal to (2/3)1/2 tImes the octahedral shear
stress) distribution in the quadrant of a plane through a cylindrical specimen of
radius R and half-height H with a length/diameter ratio of2 and subject to unit
axial ((J 1) and radial ((J 3) compression (after Balla 1960).
\
\ \ 0 1· 20 I I:;'
/ I 0
\, 0 ·132 /
\ ......

\~'
\ "'~ : I
......;.: ... : .... .. .... :: .... -.. :: : .....
~ ".::~.:,~::: :'.::::':' :' ~ ~ .:...:: ~ :
\ ............
. :: .... ::..
. I
\ • •• J
\
\

Figure 3.2 Contours of tensile stress (tension positive) computed from the
McLintock- Walsh equation (equation (3.15)) by Hawkes and Mellor (1970)
for an unconfined cylindrical specimen subjected to unit axial compression stress
and zero confining pressure. Areas of peak tensile stress are stippled.
Rock Deformation 63

variable, then following failure it should be possible to monitor the


force-resisting characteristics of the collapsing material.
Hudson et al. (1972) discuss the development of machines capable
of displacement controlled deformation, and suggest two ap-
proaches:
(a) construct machines with a high longitudinal stiffness;
(b) use a closed-loop servo-controlled testing system.
The concept of stiff testing and the basis of testing machine design,
whether it be machine stiffness or speed of response of a servo-system,
can be explained simply by reference to Fig. 3.3. This represents the
respective deformation characteristics of the testing machine and

Lowering under
compresSIon force

FORCE (CO PRESSION I

I
~
--- - - - - -
machIne R I
Iml I

ospeomef'{ :
I SI
~. ~. :
I

nm 1.5
DISPLACEMENT

Figure 3.3 Linear force-displacement curves in a testing machine system


represented by two rigid supports Rand R' containing Hookean springs to
represent machine (m) and specimen (s) stiffness. Lowering of support R
compresses both the machine and specimen, giving displacements relative to the
midpoint 0 which are positive in the case of R' and negative in the case of R
(after Hudson et at. 1972).

specimen represented by two springs between two rigid platens. As


the upper platen moves, the force-displacement relations for the
machine and specimen (assuming both springs are linear and
displacement is taken relative to the midpoint) will be given by:
(3.1 )
where K m, Ks are the machine and specimen longitudinal stiffness and
<5 m , <5s the machine and specimen displacement relative to the
64 Engineering Behaviour of Rocks

If the force-displacement characteristic is nonlinear, then equation


(3.1) becomes:
(3.2)
Salamon (1970) argues that the machine-specimen system will be
in stable equilibrium if the machine is unable to induce further
displacement in the specimen without a further input of energy from
the machine. This can be stated in terms of the energy required to
induce an elemental displacement Ab to brn and bs. Then if AWrn , AW.
are the incremental energy inputs, conditions for stable equilibrium
will be:

or

or
(3.3)
where f'(b s ) is the derivative of the specimen force--displacement
curve, positive before peak stress and negative after peak stress.
Thus since Krn is always positive, the machine-specimen system
will only be unstable whenf'(c5 s ) is negative and has a magnitude of
If' (c5s )1 ~ Krn on the unloading side of the curve (Fig. 3.4). If
1f'(c5s )1 < K rn , post-peak stress deformation will be controlled and
displacement will only be increased iffurther energy is supplied to the

•,
-t'lb »K:>
"-.. - - -
,'---_ _ ---=-' _ m
STRAIN

Figure 3.4 Typical stable ( - ) and unstable (---) loading curves in a


strain-controlled compression test system (after Hudson et al. 1972).
Rock Deformation 65

specimen. A testing machine is 'stiff' therefore provided that


Km> If'(<>s)1 whenf'(<>s)<O.
Wawersik (1968) has proposed a classification based on energy
input and output from a specimen. However, detection of this is
beyond the sensitivity of most recording equipment.

3.2 Rock deformation in compression

If a simple unconfined test is carried out by loading a cylindrical


specimen axially in compression, the resultant axial stress-strain
curve will depend on the type of rock under test. In weaker rocks there
will be distinct anelastic characteristics (Fig. 3.5), and in the stronger
rocks the curve will be nearly linear.

a STRONG ROCK(GRANI TE )

MEDIUM STRENGTH
(SANDSI0 E)

WEAl( ROCK
(SHALE )

E AXI AL

Figure 3.5 Typical axial stress-strain curves for cylindrical rock specimens
loaded axially in compression.

If, however, strain gauges are attached to the specimen circumfer-


entially (Fig. 3.6) to measure lateral strain, then even in the rocks with
a linear axial stress-strain characteristic there will be a distinct
nonlinear stress-lateral strain characteristic which would not occur
in an ideal material. This can be expressed in terms of change in
volume: after an initial compression of the specimen, there is an
increase in volume or dilation of the specimen which starts at about
40% of the peak stress and continues past peak stress. A typical
stress-volumetric strain (dilation) curve is shown in Fig. 3.7 with the
deformation stages initially proposed by Brace et al. (1966).
The explanation of this dilation is the key to understanding the
66 Engineering Behaviour of Rocks

I
. ~
CJ

\ I
\ I
\
\ DIL ATIO

~
\

- ve
E LATERAL

Figure 3.6 Typical axial (cA) and lateral (cL) strains in a cylindrical
specimen of strong rock loaded axially in compression, illustrating the effect of
fracture dilation on the lateral strain at higher stresses.

mechanics of rock deformation. The dilation of the specimen means


that space is created during the later stages of the deformation
process. Obert and Duvall (1957) also noticed another
phenomenon - during compression, small noises were emitted from
the rock sample. These noises or microseismic events were attributed
by Obert and Duvall to cracking of the rock. These cracks or
micro fractures initiated or propagated by the stress regime in the rock

a
I ill
------+---~~--/---
.I
I
J

.,
Figure 3.7 Typical axial stress-volumetric strain curve for a cylindrical
specimen of strong rock loaded axially in compression. The stages indicate : I,
linear recoverable deformation; II, microfracture initiation and propagation
through the specimen at a controlled rate; I II, rapid crack propagation leading
to eventual disintegration of the specimen.
Rock Deformation 67

were demonstrated by Cook (1965) to increase with increasing load


(Fig. 3.8). They have been treated rhetorically and theoretically by
various workers but some of the most important experimental work is
by Scholz (1968a,b) working in Brace's laboratory, who set out to see
first of all if dilation could be accounted for solely by micro-
fracturing, and then to see whether the phenomenon fitted accepted
theoretical concepts of brittle rock fracture and if it played a
significant role in deformation of more ductile rocks.
Scholz's experiments and experimental techniques are described in
the literature. Six rocks (Table 3.1) were selected for study and 16 mm
diameter, 50 mm long specimens were subjected to loading in a stiff
(10 6 kN mm - 1) test rig at a constant strain rate of 10 - 5 S - 1.
Volumetric strains were measured with strain gauges, and micro-
seismic events in the frequency range 100 Hz-1 MHz were measured
using barium titanate transducers attached to the specimen.
Typical frequencies of microseismic events for the two rocks with
the most widely varying physical properties are illustrated in Fig. 3.9.
There is close similarity in behaviour both together and with Cook's
results. Initially, there is a flurry of activity at low stresses, attributed
to closing of existing cracks and pores, and roughly proportional to
the porosity of the rock. This is follwed by a very low level of activity
up to about 40% of peak stress. There is then a steady increase in
activity until just before fracture when there is a rapid increase in
activity.
Comparison of accumulated frequencies of microseismic events
and volumetric strains for granite in Fig. 3.10 indicates a close
correlation between micro fracturing and dilation. This was found
by Scholz to be the case over the dilatant region for all the rocks
tested, the correlation being well within experimental error except in
the region above 95 % of the peak stress, where micro fracturing
activity was found to accelerate rapidly. In Fig. 3.11 data normalized
with respect to peak stress are reproduced for the five rocks tested.
Although the rocks had unconfined strengths varying from
50 MN m - 2 to a very high 505 MN m - 2, the behaviour of all the
rocks was similar when loaded uniaxially in unconfined compression.
It can be concluded therefore that during rock deformation-
leading to breakdown - fractures are created in the rock, evidence for
which is available both in the form of dilation and microseismic
emissions. Thus to understand better the mechanism of breakdown in
rock, it is necessary to look at the factors affecting propagation of
microfractures within its structure.
150
N
St CLOUD
E GRA ITE
~100
::E
V'>
~ 50
a:
l-
v>

o 0-1 02 0-3 04 0'5 Q6 0·7 0·8 STRAI %

o 0.1 0.2 03 O.G 0.5 0.6 0.7 O. B STR AIN %

150
N TENNESSEE
E
MARBLE
%100
::E

~
w
~ 50
'"
3 STRAIN%

STRAIN %

Figure 3.8 Stress-axial strain curves for cylindrical specimens of St Cloud


Granite and Tennessee Marble, loaded to fracture in uniaxial compression,
alongside envelopes indicating the frequency of microseismic activity during
deformation (after Cook 1965).
Rock Deformation 69
Table 3.1 DescriptIOn of rock used by Scholz (l968a)

Average Uniaxial
Unit particle compressive
weight Porosity diameter strength
Rock (kNm- 3 ) (%) (mm) (MN m- 2 )

Westerly Granite 26.5 0.9 0.85 281


Rutland Quartzite 26.4 0.5 0.30 505
Marble 26.9 1.3 0.20 50
Colorado Rhyolite Tuff 17.7 41 0.01 91
Pottsville Sandstone 25.4 3.0 0.20 230
San Marcos Gabbro 29.7 0.2 1.0 215

3.3 Mechanics of microfracture

Rocks comprise on a micro scale agglomerates of mechanically


bonded particles containing numerous flaws, pore spaces and
microcracks. These can occur at grain boundaries or in the mineral
grains themselves. A classification of microcracks based principally
on genesis and adapted from the work of Simmons and of Richter is
included in Attewell and Farmer (1976). The description of rock
deformation phenomena in the previous section gives several clues to

(0) (b)

If vs. EA
";-
...
200
COLORADO 40 ~
WESTERLY 0-

-
N
GRANITE
RHYO LITE z
w
E
TUFF
z
L
...~
0
>-
b 100
20 ~
Vl
....
Vl ....
::>
cr:
0- ....
0
cr:
Vl .....

0 0 ·2 0· " 0 ·1 u·L 0 ·6
AX IAL STR AIN (A % AXIAL STRAI [ A%

Figure 3.9 Plots of stress against axial strain and frequency of microseismic
events during uniaxial compression of specimens of Westerly Granite and
Colorado Rhyolite Tuff (after Scholz 1968a).
70 Engineering Behaviour of Rocks

~
:z:
0-2
;;: WESTERLY
a: GRA ITE
l-
V>
wO·15
0::
I-

1:
s! 0-1
=>

'='
.....
V!

;j0 - 05-1-------~
:::=

O~==~~~------~--~----~
50 fIJ 70 eo 90 100
PERCENT OF PEAK STRESS

Figure 3.10 Plots of inelastic volumetric strain (the difference between total
volumetric strain and computed 'elastic' volumetric strain of BA - 2B L ) against
percentage of peak or fracture stress for three unconfined compression tests on
Westerly Granite. Included on the plot are points determined by Scholz (1968a)
to represent accumulated frequency of seismic events.

the mechanical description of microfracturing. Two are of particular


importance:
(a) The microfractures are initiated at a stress level related to the
peak stress or 'strength' of the rock.
(b) The dilation in a direction normal to the major (compressive)
principal stress indicates that the microfractures spread in a

~
z
GABBRO
MARBLE
I
;;:
'"tii 0 ·1
TUFF
SANDSTONE
QUARTZITE ._- - ---
/

so 60 70 80 90 llO
PERCENT OF PEAK STRESS

Figure 3.11 Plots ofinelastic volumetric strain (see Fig. 3.1 0 caption) against
percentage offracture stress and a dimensionless representation ofaccumulated
frequenc y for five rocks (after Scholz 1968a).
Rock Deformation 71

direction parallel to the major principal stress - in other words


they are probably a result of tensile failure.
These two observations fit well with the best mechanical descrip-
tion of brittle fracture initiation developed to date, namely that of
Griffith (1921, 1924) originally proposed to explain the fracture of
glass in tension and modified by Orowan (1949) and Brace (1960),
among others, for the case of rocks loaded in compression.
Griffith's analysis is based on the assumption that cracks can be
represented in two dimensions by an ellipse (Fig. 3.12), and postulates
that:
(a) Cracks are initiated when the tensile stress at at the tip of an
elliptical crack is sufficiently high to provide the energy required

',, ~ bo
">
/ tTJ

/ ~ab

Figure 3.12 Elliptical crack in a biaxial stress field.

to create new crack surface. This can take the form for an elastic
material:
a = (2YE)1 /2 (3.4 )
t nC
where E is the modulus of elasticity, Y is the specific surface energy
required to satisfy a unit of crack surface area and C is the crack
half-length.
(b) Although the cracks may be randomly orientated (this is rarely
the case in rocks, which tend to be anisotropic), the direction of
crack extension will be normal to the direction of maximum
tensile stress in the specimen.
72 Engineering Behaviour of Rocks

Equation (3.4) can be obtained by a simple energy balance (a


detailed solution is given by Bieniawski (1967)) equating the energy
input to the rock to the energy stored as elastic strain energy, and the
surface energy required to satisfy the surfaces of newly formed cracks.
The derivation of Griffith's criterion for the spreading of a
randomly orientated elliptical crack in a two-dimensional stress field
is given in Jaeger and Cook (1969) and is worth considering in some
detail. The stress field (Fig. 3.12) can be resolved into a normal and
shear component aa and tab' if it is assumed that the stress acting along
the major axis of the elliptical crack, ab , will have negligible influence
on the stress at its tip. a, b are the minor and major axes of the ellipse.
a a and tab can be related to a 1 and a 3 using the Mohr representation
of stress (see Fig. 2.7):
(3.5)

(3.6)

and the tangential stress at on the crack boundary can be shown to be


given in terms of aa and tab by:
2aa[sinh 2Ro +cos 2T exp(2Ro) -1] + 2tab sin 2T exp(2Ro)
at =
cosh 2Ro - cos 2 T
(3.7)
where Rand T are the elliptical coordinates, so that in Cartesian
terms:
x=c sinh R sin T, z=c cosh R cos T
Ro is the intercept of R on the ellipse boundary and c is a complex
constant which is equal to half the major axis when R = R o, or half the
crack length as in equation (3.4).
If Ro is small, as in the case of a flat crack, and T is small
(corresponding to the tips of the ellipse) then the tangential stress at
the tips of the ellipse can be obtained from equation (3.7):
2(aaRO +'tab T )
a=-~'---------;;;::-- (3.8)
t R6+ T2
Maximum and minimum values of at can be obtained from
dat/dT=O given by differentiating equation (3.8):
(3.9)
where the negative sign gives the maximum tensile stress at the tip of
Rock Deformation 73

the ellipse. Then by substitution from equations (3.5) and (3.6):


ROatmax=t[(al +a 3 )+(a 1 -a 3 )cos 28]
- t[2(ai + aD + 2(ai - a~)cos 28]1/2 (3.10)
The critical angle 8ent between the major axis of the ellipse and the
direction of the major principal stress at which at max has a maximum
value (a~max) will be given by da t max/d8 =0, so by differentiating
equation (3.7):
(3.11 )

Thus by combining equations (3.10) and (3.11) a general criterion


for maximum tangential (tensile) stress at the tip of the ellipse (and the
crack it represents) can be obtained in the form:

R * _ 1 (0'1 -0'3)2
OatmaX--4(
0'1
+ 0'3 ) (3.12)

This can be extended to a general criterion for failure if some


material characteristic can be substituted for Roa~max in equation
(3.12). The most simple approach is to go directly to equation (3.7),
where it can be shown that if 8=90°, or the crack is parallel to the
major principal stress, then the tensile stress on the crack surface will
have a maximum value:
(3.13 )
If 0'3 is equated to the tensile strength of the material, aTf , then a
failure criterion emerges in the form:
1 (0'1 -0'3)2
aTf =8 (3.14 )
(0'1 +0'3)

provided that the condition for a negative (tensile) 0'3 in equation


(3.11), 0'1 + 30'3 < 0, obtains. This is the basic statement of the Griffith
criterion of failure in a biaxial stress field. Extension to a three-
dimensional stress field was considered by Murrell (1963).
In deriving equation (3.14) it is assumed that the crack retains. its
shape to the moment of failure or extension. This is not necessarily the
case, and McLintock and Walsh (1963) modified the Griffith criterion
to allow for friction along the surface of closing cracks:
aTf=!a 1[(tan 2¢ + 1)1/2 -tan ¢] -!1T3[(tan2¢ + 1)1/2 +tan ¢] (3.15)
where tan ¢ is the coefficient of internal friction of the rock.
74 Engineering Behaviour of Rocks

It should be stated that, whereas the Griffith criterion illustrates the


mechanism of rock failure, it cannot describe or predict it accurately.
This is because the inhomogeneity intrinsic in most rocks will
produce fluctuations in the stress field which will both limit the
propagation of cracks after initiation and reduce the expected
fracture strength of the specimen as a whole. Nevertheless it does
provide a tool with which to examine experimental data and to
develop an understanding of the processes which lead to rock
fracture. For instance, both equations (3.14) and (3.15) provide
expressions for the maximum tensile stress at the surface of a crack.
Contours of either the Griffith or the McLintock-Walsh tensile stress
parameter in a rock specimen will allow points at which the rock may
fracture to be isolated and explain observed breakdown phenomena.
Even so, the process is further complicated by the presence of plastic
minerals. Tapponier and Brace (1976) state that:
... minerals which amount to only 5-10% by volume may limit
strength in brittle rock, if they (a) are capable of plastic slip or (b)
have unusually high or unusually low elastic modulus relative to
other minerals in the rock.
Examples of plastic and other mechanisms of crack formation
observed in Tennessee Marble are given in Fig. 3.13.

3.4 Rock macrofracture

Hawkes and Mellor (1970) took equation (3.15) and used it to

DIRECTION OF
COMPRESSION
'Yt
Ao GUOE LAMELLA
B=GRAI BOUNDARY
&..
A
C

I
C=CRACK

Figure 3.13 Schematic diagram (after Olsson and Peng 1976) showing some
of the crack initiation mechanisms observed in Tennessee Marble. Maximum
compression is in the direction indicated.
Rock Deformation 75

calculate, from Balla's (1960) data (Fig. 3.1), the McLintock-Walsh


tensile stress parameter for a cylindrical specimen under uniaxial
compression and restrained at the sample ends. The contours
reproduced in Fig. 3.2 illustrate that the most critically stressed zones
are at the specimen centre and at the edge of the platen-specimen
interface. The latter effect may be reduced by using elongated
specimens and low friction or composite end platens.
Scholz (1968b), using an end platen system designed by Mogi
(1966a), examined the location of signals emitted from a 25 mm
diameter by 100 mm long specimen of Westerly Granite during
uniaxial compression. The data from Fig. 3.10 have already been
discussed briefly. Scholz showed that at stresses up to 95% of
compressive strength there was a relatively stable relationship
between the accumulated frequency of micro fracturing, N, and stress,
taking the form of a normal distribution (Fig. 3.14) and implying that
the local stress producing each event was an independent random
variable. Above the 95 % strength level N was related to stress and a
rapid acceleration in dilation or micro fracturing occurred.
The location of 22 events to within 3 mm in a test specimen is
illustrated in Fig. 3.15 which shows the central 75 mm of the specimen
(between resin collars) and also the trace of the fracture plane which
eventually led to fracture. Fig. 3.l5(a) shows events below 95% of

-.e
z
:;;: ~

IX
t;;o ., 1-0 Z
u >-
a: ORMALI (j ,K] aI • u
....
Z
t;:; =>
I: OISTRI BUTIO 0
=> w
IX
---'
C) u..
> C
u "-'
tiO,OS 0.5 3
5
.... =>
I:
;;; ::>
u
u
<0:

50 60 70 80 90 100
PERCENT OF PEAK STRESS

Figure 3.14 Fit of a normal distribution with mean ii and standard deviation
K2ii to the data on accumulated frequency in Fig. 3.10 (after Scholz J968a).
76 Engineering Behaviour of Rocks

~
Q
~. ~,
b
X

~/
I

9/ 0

/
I
~
/
/ \
\
\
~
, ~_/ /
/
I

/
\OJ
\
\
'--
/
/

Q = STATIC b = OYNAMI (

Figure 3.15 Front, top and side views of the central section of a test specimen
of Westerly Granite showing the location of microseismic events which occurred
(a) in the 'static' cracking regime below 95 %offracture or peak stress and (b)
in the dynamic cracking regime above 95% of fracture stress. The ultimate
fracture plane is traced by the broken line (after Scholz 1968b).

compressive strength; Fig. 3.15(b) events above it. Allowing for a


certain amount of bias produced by the position of the transducers,
the results support Scholz's proposition that at low stresses events
occur randomly through the specimen and are not closely related to
the ultimate fracture plane, although they tend to fall in the specimen
regions with the highest tensile stress parameter. At high stresses,
however, the events tend to cluster and are closely associated both
with the fracture plane and with the regions of highest tensile stress.
Thus at a critical level of stress, crack density or dilation, there is
evidence of coalescence of cracks to form a fracture plane - or
possibly in 'ductile' rocks a series offractures -leading to breakdown
of the rock, and associated with a rapid increase in seismic activity.
This is the basis for seismic monitoring of active faults to predict
earthquakes, and of highly stressed zones in mines to predict
rock bursts.
Rock Deformation 77

FAULTED PEAK
STRENGTH

RESIDUAL
FORHATlO OF STRE GTH
SECOND HACROCRACK

a. A~RAGE AXIAL STRESS


''Yv I'OLUHETRIC STRAIN
C. A ERAGE AXIAL STRAI
f FR[QU ENty OF SIESHIC
E ENTS

Figure 3.16 A description of rock deformation (after Price 1979).

3.5 The complete rock deformation curve

The essential stages of deformation of a cylindrical rock specimen


tested in uniaxial compression are outlined in Fig. 3.7. The descrip-
tion of the mechanics of crack propagation and coalescence to form
the shear (or conjugate shear) plane* of Fig. 3.15 allows these stages
to be extended to give a complete description of rock deformation in
Fig. 3.16. This includes six stages which are indicated by the letters
A-F in Fig. 3.16:
*It can be argued that use of a spherically seated platen In testing, allowing some
rotatIOn following Imtlal shear, will tend to form smgle shear planes Plane seated
platens will tend to be associated with conjugate shears.
78 Engineering Behaviour of Rocks

(a) Stage I (A) - When the rock is initially stressed any pre-existing
microcracks or pore space orientated at suitable angles to the
applied stress will close. This causes, in weaker and more porous
rocks, an initial nonlinearity of the axial stress-strain curve.
(b) Stage II (B) - The rock has a near-linear axial and lateral
stress- strain curve which is largely recoverable. This is accom-
panied by compression - again recoverable - which has distinct
similarities with elastic compression, although the Poisson's
ratio, particularly in stiffer unconfined rocks, tends to be low. It is
quite reasonable to describe the deformation characteristics as
elastic in this stage. There is a minimum of seismic activity during

Figure 3.17 Sections through 75 mm diameter cylindrical specimens of


Portland Stone where a test has been stopped at or about peak stress. The start
of the shear planes from points of peak tensile stress (see Fig. 3.2) in the
specimen can clearly be seen.
Rock Deformation 79

this stage and it is also reasonable to argue that micro crack


propagation only starts at the upper boundary of about 35-40%
peak stress.
(c) Stage III (C) - The stage is characterized by the onset of
dilation and by a near-linear increase in volume, which is offset
against continuing compression. There is also a near-linear axial
stress-strain curve which like Stage II is nearly fully recoverable. It
can be proposed that microcrack propagation occurs in a stable
manner during this stage and that microcracking events occur
independently of each other and are distributed throughout the
specimen. The upper boundary of the stage is the point of
maximum compaction and zero volume change. It occurs at
about 80% peak stress and has occasionally been used as a
reference point for critical state descriptions of rock deformation
(see Chapter 4).
(d) Stage IV (D) - This stage is characterized by a rapid acceleration
of microcracking events and of volume increase. The spreading of
microcracks is no longer independent and clusters of cracks in the

Figure 3.18 A 75 mm cylindrical specimen of Portland Stone deformed in


triaxial compression to 10% axial strain.
80 Engineering Behaviour of Rocks

zones of highest tensile stress (Fig. 3.2) tend to coalesce and start
to form tensile fractures or shear planes - depending on the
strength (and degree of confinement) of the rock.
(e) Stage V (E) - This is the stage where the rock has passed peak
stress, but is still intact, even though the internal structure is
highly disrupted. In this stage the crack arrays fork and coalesce
into macrocracks or faults. It is possible to talk about failure at
this point; more sensible to talk about strain softening deform-
ation. In this description, at peak stress the test specimen starts to
become weaker with increasing strain. Thus further strain will be
concentrated on weaker elements of the rock which have already
been subjected to strain. This in turn will lead to zones of
concentrated strain or shear planes as in Fig. 3.15.
(f) Stage VI (F) - In this stage the rock has essentially parted to
form a series of blocks rather than an intact structure. These
blocks slide across each other and the predominant deformation
mechanism is friction between the sliding blocks. Secondary
fractures may occur due to differential shearing. The axial stress
or force acting on the specimen tends to fall to a constant residual
strength value, equivalent to the frictional resistance ofthe sliding
blocks.
The condition of the rock in two of the later stages can be illustrated
in Figs 3.17 and 3.18 which show test specimens of Portland Stone.
The specimens in Fig. 3.17 have been obtained from tests which were
stopped at the peak stress at the end of Stage IV. Sections through the
specimens show clearly the onset of shear planes starting at the highly
stressed corners of the specimens. In Fig. 3.18, the test, albeit with a
degree of confinement, has continued to 10% axial strain and it can be
seen that the specimen comprises a series of quite discrete blocks.
4 Rock Strength
and Yield

The concept of rock strength or strength failure as a result of


laboratory testing is difficult and complex to define. Stated most
simply, rock strength is the peak stress in a uniaxial or triaxial
compression deformation process. During this process there is no
unique point at which rocks fail or collapse. The same is equally true
in the vicinity of most structures in rock. A hole in the ground
represents a small opening in a large continuum. Energy released by
fracturing at a part of the structure is rapidly accommodated in the
continuum. Only in the case of isolated pillars or the edges of plates in
the earth's crustal structure do the conditions exist for rapid release of
energy - in exceptional forms such as rock bursts or earthquakes.
In order to understand strength, it is necessary to consider the later
stages of rock breakdown illustrated in Figs 3.16--3.18 and to explain
how these may be accommodated in a strength criterion, capable of
use in design in rocks around underground openings.

4.1 Rock strength criteria

Three criteria to describe the strength of rocks are usually quoted:


(a) The Coulomb criterion
(4.1)
where!f and anf are the shear and normal stresses across a plane (see
Fig. 2.6) at the point at which strength failure takes place. In the
context of Fig. 3.16, this will be the peak stress.

81
82 Engineering Behaviour of Rocks

(b) The Mohr criterion


(4.2)

which is the general case of the straight line indicated by the


Coulomb criterion, and implies that the strength envelope
represented is curved - usually concave downwards. This allows
for a degree of plasticity or yield at higher confining pressures.
(c) The modified Griffith or McLintock-Walsh criterion
(4.3)
which is a linear form derived from equation (3.15). Alternatively,
a curvilinear form similar to the Mohr envelope can be derived
from equation (3.14) in the form
(4.4)

having an abscissa intercept of (J Tf and an ordinate intercept of


2(JTf'

The most common assumption is some form of straight line, and


this is acceptable at engineering stresses in most rocks. Inherent in all
the approaches is the assumption that since !f is the driving shear
force causing strength failure, the intermediate principal stress is not
an important factor. This is demonstrably not the case with stresses
on a plane (see Fig. 6.10).
The Coulomb criterion is illustrated in Fig. 4.1. It is possible,
because of the linear form, to develop notional relationships between
shear, compressive and tensile strengths. For instance, substituting
for! and (In in equation (4.1) from equations (2.28) and (2.29) gives a
restatement of the strength criterion in the form (where tan ¢ = /1):
(Jlf[(/12 + 1)1/2 -/1] -(J3f[(/12 + 1)1/2 +/1] =2c (4.5)
similar to equation (3.15).
Then if (J 1 = 0, (J 3 = - (J Tf (tensile strength), and
(JTf[(/12 + 1 )1/2 + /1] = 2c (4.6)

and if (JI =(Jcf (compressive strength), (J3=0, and


(Jcf(/12 + 1)1/2 - /1 = 2c (4.7)

whence
(J Tf (/12 + 1)1/2 - /1
(4.8)
(J cf (/12 + 1)1/2 + /1
Rock Strength and Yield 83

ern

90-8

0,

Figure 4.1 Coulomb failure criterion ' f = c + (in ( tan cp showing the position
of shear fracture plane angles in a Mohr circle constructed to touch the strength
envelope.

and

~-~=1 (4.9)
a cr aTf

These can be used as a very simple basis for rock classification (see
Hoek 1966).
The r-a construction of Fig. 4.1 is simple and allows ready
estimation - even with an empirical Mohr envelope - of the angle f)
of a potential failure plane through the normal to the envelope
passing through 1(a\ +(3) on the an axis. It is however sometimes
more convenient to express the data in qr, Pr space or a If' a 3r space
(Fig. 4.2).
In these cases the linear relation may be expressed as:
(4.10)
T
1
MN /m

100 / / - ----z__ ~
/~---

50

Qf: 0"11- OJ
Pf: O"lf .20"Jf
--3-

50

25 50 75 100 P f MN/m 2

50

25 50

Figure 4.2 Alternative methods by which the peak strength envelope for
Portland Stone may be plotted from the data in Fig. 4.15. Note that values of ¢
computed from the qf, pf and a If' a3f plots are higher than from the T, an plot.
Rock Strength and Yield 85

where it can be shown that M = (6 sin ¢ )/(3 - sin ¢), or as:


(4.11 )

where Kp = (i + sin ~)/( j - sin ~) Is sometimes called the triaxial


stress factor or flow factor and is numerically identical to the
coefficient of passive earth pressure and the bearing capacity factor,
N <1>' of foundation engineering.

4.2 Yield criteria

The Mohr and Griffith criteria with a reducing slope on a r - O'n plot
are essentially yield criteria - the implication being that with
increasing confining pressure there is a reduced increase in deviatoric
stress. If the confining pressure is increased sufficiently - possibly
with an increase in pressure and reduction in strain rate - it is logical
to assume that the deviator stress required to cause failure will remain
constant and independent of any increase in confining pressure.
It is necessary to enlarge upon this statement, by defining as far as
possible the term failure in rock testing. At low confining pressures
and in strong rocks, failure will be recognizable as a shear plane or
tensile crack accompanying, on a strain-softening stress-strain curve,
an easily identifiable peak stress. At higher confining pressures, failure
is less easily recognized as excessive deformation (or strain) on a
strain-hardening curve. The two types of curve are illustrated in
Fig. 4.3.
The basic mechanics of strain-softening deformation have been
discussed in Chapter 3, and both types of deformation show a strong
analogy with soil deformation (see Atkinson and Bransby 1978).
In the strain-softening case, the specimen will tend, approaching a
peak stress, to become weaker with increasing strain. Thus strain will
tend to be inhomogeneous and further strain will be concentrated in
the weaker elements of the rock which have already been subjected to
the most strain. Thus following peak stress, these zones of concen-
trated strain or shear planes will develop as in Fig. 3.18. In the case of
strain-hardening deformation, specimens of rock become stronger as
they deform. Consequently, strain will tend towards homogeneity
throughout the confined specimen, since those elements of the rock
which have strained most will be stronger than those which have
strained less. Then at failure, in this case defined by large axial strain,
the specimen will be uniformly deformed and cataclastically
86 Engineering Behaviour of Rocks

INCREASltIj STRENGTH WITH


INCREASING STRAIN

STRAIN HARDENING

0,

STRAIN SOFTENING

RESIDUAL STRENGTH

Figure 4.3 Typical strain-softening and strain-hardening types or rock


deformation. The former is typical of high-strength rocks at low confining
pressures, the latter of low-strength rocks.

fractured - probably with some trace of Liiders lines to denote some


shear as in Fig. 4.4. Provided the specimen continues to deform
homogeneously, it will ultimately reach a state at which large
distortions will occur without change in volume. This ultimate state
of plastic yielding is known as the critical state and is the basis of
critical state soil mechanics (see Atkinson and Bransby 1978;
Schofield and Wroth 1968).
The critical state should not be confused with transition between
strain-softening and strain-hardening behaviour which is effectively a
brittle-ductile transition beyond which some dilation may occur, but
where deformation has strongly ductile characteristics. This can be
illustrated by data from Scholz (1968a) on marble in Fig. 4.5, which
illustrate several important effects of increasing confining pressure:
(a) Confining pressure increases the strength of the rock and the
degree of post-yield axial strain hardening. These effects diminish
with increasing confining pressure.
(b) At low confining pressures there is increasing dilation, which
reduces at higher confining pressures until at the highest,
400 MN m - 2, there is little or no dilation.
Rock Strength and Yield 87

Figure 4.4 A specimen oj saccharoidal limestone subjected to large axial


strain at high confining pressure (see Jurther data in Fig. 4.16) and exhibiting
signs of plastic deformation as well as multiple shear planes or Liiders lines.

(c) If, as in Fig. 4.6 (in this case at an arbitrary 1 % axial strain),
Poisson's ratio, axial stress and the slope ofthe stress-axial strain
and stress-volumetric strain curves are plotted, it can be seen that
they approach a limiting value with increasing confining
pressure. The limiting value of Poisson's ratio is 0.5.
This indicates that, at a confining pressure around 400 MN m - 2,
something similar to plastic flow develops. Scholz quotes from Hill
(1950) the requirements for plastic flow in a polycrystalline material:
... it is unaffected by confining pressure, provides little or no source
for radiative elastic energy and produces no volumetric strain.
This will be expanded in Section 4.6.
88 Engineering Behaviour of Rocks

300
[0 FI I (j
PRESSURES
IN I'N/m2

z
<
'" 0·4
u

~~O.8
~ >
:::»
c5 <l
>

Figure 4.5 Axial stress-axial strain and volumetric strain- axial strain curves
for marble loaded in triaxial compression at confining pressures up to
400 MN m- 2 (after Scholz J968a).

There are various methods for describing the yield behaviour


threshold of rock. The most simple have been developed to describe
yield in metals and two may be mentioned briefly. One ofthese is the
Tresca criterion which states that plastic deformation occurs in a
material when the maximum deviator stress reaches a critical value:
(4.12)
or as modified by Bishop (1972) to allow for some increase in yield
resistance, due to friction, with increased confining pressure at some
point before the critical state:
(4.13)
The other yield criterion is ascribed to Von Mises or Von
Mises-Huber-Hencky and states that plastic deformation occurs in
a material when the distortional strain energy reaches a critical value:
(a 2 - 0'3)2 + (a 3 - a d2 + (a 1 - 0'2)2 = 2O'~ (4.14 )
or in modified form:
(a; - O'~)2 + (a; - O'~)2 + (O'~ - 0';)2 =~(X2(O'~ + a; + O'~f (4.15)
In each case (X is a variable parameter. Bishop shows that neither of
these yield criteria in their original or extended forms fits the data for
rock or soils with any degree of accuracy, and a more satisfactory
Rock Strength and Yield 89

N
e dO'"/d(
;Z 250 0
~5 5
1:
(J' (,
2 4
z
;:{ N
POISSON's 3 E
....
0::
<II RATIO ' ·5 z

- , '"
')I
~
2 w
ca ~
:r b
>-
'"
~ 0·5 '
'l:>

0::
.....
<II

tlO 200 llO 400


(0 FINING PRESSURE MN 1m2

Figure 4.6 Strength at 1 %axial strain, Poisson's ratio and the slope oj the
axial stress-axial strain curves from Fig. 4.5 plotted against cotifining pressure
(after Scholz 1968a).

model is almost certainly the critical state model developed for soils.
This introduces the quantity of specific volume of a unit volume of
solids (equal to 1 +e, where e is the voids ratio) and is based on the
premise that for all particulate materials there will be a unique
combination of specific volume, confining pressure and deviatoric
stress at which a specimen under test will continue to yield without
further change in volume. This idealized plastic state is called the
critical state and was proposed by Roscoe et at. (1958) as a basis of
reference for the mechanical description of soils.

4.3 The critical state concept

Schofield and Wroth (1968) developed two models to describe the


yielding of soil as an isotropic rigid-plastic and elastic-plastic
material. These models describe a series of yield surfaces in principal
stress space, which in addition to defining 'hardening' and 'softening'
of the material as a function of plastic distortion, also define the
plastic volume change as a function of spherical stress. The particular
aim of the model is to describe a particulate material mechanically
rather than empirically - as in the conventional approaches t<.1 soil
mechanics. In this way it is possible to approach some fundamental
'properties' of soil rather than the results of arbitrary laboratory
experiments. It is worth describing the model in order to see if the
same approach may be valid for rocks.
90 Engineering Behaviour of Rocks

The critical state model proposed by Schofield and Wroth (1968)


and Atkinson and Bransby (1978) comprises a series of yield surfaces
in q, p space corresponding to different values of specific volume, v.
Fig. 4.7 illustrates the model in which two main yield surfaces can be
identified, the Rvorslev and Roscoe surfaces, separated by the critical
state line. These surfaces are also termed 'state boundaries', since they
separate the states which any particular specimen can achieve from
those which it can never achieve. Thus, the behaviour of a particular

q'

I
P
(RITI(AL STATE LINE

NORMAL CONSOLIDATIO N LINE

Figure 4.7 A critical state modelfor soils (after Atkinson and Bransby 1978).
Rock Strength and Yield 91

material tested triaxially can be described or determined using such a


model.
The critical state line is defined as the single and unique line of
failure points for both drained and undrained tests, where failure is
taken to be the state at which large shear distortions occur with no
change in stress or in specific volume, The position of the critical state
line in q', v, pi space is defined by the equations:
ql=Mp' (4.16)
and
V=r'-A In pi (4.17)
where q' is the effective deviator stress, pi the effective spherical stress,
v the specific volume, M and A are constants, and r ' is the specific
volume of the soil at the critical state with pi = 1.0 kN m - 2.
The Roscoe surface is the state boundary surface for normally
consolidated and lightly overconsolidated soils. The idealized stress
paths (Fig. 4.8(a)) in q, pi space for specimens of these soils during

CRITICA L
STATE
LINE

ROSCOE
SURFACE

~------------~--~~~-.- p

la)
q

CRITICAL
STATE
LINE
HVORSLE
SURFACE

p
( bl

Figure 4.8 Loading stress paths for (a) normally consolidated or lightly
overconsolidated soils and (b) heavily overconsolidated soils or rocks.
92 Engineering Behaviour of Rocks

triaxial testing rise to the state boundary and then follow the
boundary to the critical state line and failure.
When specimens of heavily overconsolidated soil are compressed
the stress-strain curve rises to a peak and subsequently reduces to a
residual value with continued straining. The idealized stress paths in
q, pi space for such samples (Fig.4.8(b)) rise to a linear state
boundary, above the projection of the critical state line, and then
follow this boundary to the critical state line and failure. This
boundary is termed the Hvorslev surface and is bounded at one end by
the critical state line and at the other by the tension cut-offline, where
the specimen will fail in tension, as discussed in the previous chapter.
In practice, however, heavily overconsolidated soils are more likely
to fail prematurely, due to inhomogeneities in the specimen under
test. Two particular states for these soils must therefore be
distinguished. The first is the 'failure state' which occurs at the point of
maximum deviator stress. The second is the 'ultimate state' where the
conditions for the critical state line are met; that is, the state at which
large shear distortions can occur with no change in stress or volume.
Consequently, although a particular specimen may have only
attained its failure state, elements within it may have attained their
ultimate state and thus their critical state.
Inherent in the critical state approach is the requirement that the
energy input to the test specimen from the test system for any given
strain increment can be controlled, so that there will never be a
negative energy input into the specimen. This condition is known as
stable deformation. In triaxial testing this requirement can easily be
met in soils looser than the critical state, but in rocks and soils denser
than the critical state it can only be met at axial stresses less than the
peak stress or at high confining pressures. Barton (1976a),
Gerogiannopoulos and Brown (1978) and Brown and Michelis
(1978) have concentrated on the pre-peak stress curve with interest-
ing results.
It is nevertheless possible to find in the literature test data on rocks
tested at high confining pressures which fit the critical state model.
These include the tests illustrated in Fig. 4.5 by Scholz (1968a) at
confining pressures of between 25 and 400 MN m - 2 on weak marble
specimens (50 mm long by 16 mm diameter) with an average
unconfined compressive strength of 40 MN m - 2 and grain diameter
of 0.20 mm, at a strain rate of 10 - 5 S - 1 in a machine of high stiffness
(lOs MN m - 2). These deformed cataclastically, breaking down
through gradual loosening of grains into individual grain particles
Rock Strength and Yield 93

during the test. The ultimate form of the marble was like a granular
material. Except in the case of the unconfined test, deformation was
stable.
In Fig. 4.9 Scholz's results are replotted using Schofield and
Wroth's terminology as q=O' l -0'3 against pi =t(O' l +20'3) for
constant volumetric expansions. The initial void ratio was 0.013 and
the specific volume 1.013. The maximum axial strain to which the
tests were carried out was 1.5 %. Although neither the curves nor
Scholz's results are complete, they appear to form a series of yield
curves which make up the three-dimensional state boundary or
H vorslev surface discussed earlier and plotted data appear to
approach the critical state. For comparison, Barton's (1976a)
suggested critical state lines (0'1 = 30'3' drained specimens; 0'1 = 20' 3,
undrained specimens) have also been included. Mogi (1966b)
suggests 0'1 = 3.40'3 to represent a brittle-ductile transition.
The evidence for the Hvorslev surface from Scholz's data IS
sufficiently convincing to justify the critical state model as a

0,-30") ¥
~ o,ooa
150 , ~~ 0, -al,

100

p - ", I , . 20,1
HNlm'

Figure 4.9 Constant volumetric strain (Ll V/V) yield curves for marble tested
in triaxial compression calculated from data in Fig. 4.5.
94 Engineering Behaviour of Rocks

description of a rather special rock at high pressures. It is also possible


with this type of marble, which tends to uniform deformation, and
these near-tectonic pressure levels to provide some evidence for the
existence of the Roscoe surface (see Edmond and Patterson 1972) in
rock. This effectively requires that the rocks be hydrostatically
stressed to a value of p greater than the critical state. In rock
engineering, however, maximum confining pressures of the order of
40 MN m":' 2 are likely to be encountered and this may require a
modification of the critical state approach. A series of tests on
sedimentary rocks were carried out in the author's laboratory in
order to investigate deformation at lower confining pressures.

4.4 Triaxial testing

Strength criteria are invariably obtained empirically from triaxial


testing. Triaxial testing aims to simulate the conditions which may
occur in a rock material around an excavation or under a foundation
where the rock will be subjected to a confining pressure and
deviatoric stress.
The method of triaxial testing is outlined in the Suggested Methods
of the ISRM (Brown 1981) and in the relevant ASTM Designation
or British Standard. The basic equipment is the testing machine
for compression testing - which should be servo-controlled
(Section 3.1) - and a hydraulic pump, pressure intensifier or other
system sufficient to maintain a constant confining stress.
For rapid testing, to obtain a peak stress for a series of confining
pressures, a simple cell (Fig. 4.10) ofthe type developed by Hoek and
Franklin (1968) may suffice. This is available in various sizes and can
be used in the laboratory or in a portable compression rig in the field.
The method of testing is different to most triaxial tests (see Fig. 2.8) in
that a confining pressure is applied to the specimen in a stiff reusable
membrane before the axial force is applied. The membrane has the
advantage of allowing rapid testing and the disadvantage of
restricting deformation after peak stress.
A more sophisticated cell, designed by Price (1979) for tests to large
axial strains on Coal Measures rocks, is illustrated in Fig. 4.11. This
incorporates many of the desirable features needed in a simple cell for
obtaining information on the peak and residual strength characteris-
tics of rocks at moderate confining pressures.
The cell was designed to accommodate and measure large axial and
Rock Strength and Yield 95

Figure 4.10 The Hoek-Franklin type of triaxial eellfor simple triaxial testing
of rock.

volumetric strains. Based on Crouch's (1970) modification of a


dilatometer used by Bridgeman (1949), the principal feature was a
pressure relief valve for maintaining a constant cell confining pressure
in the range 3.5-70 MN m -2. The valve was mounted in series with a
pressure gauge and the cell, and the volumetric strain of the specimen
was computed, as a percentage, from the volume of oil Vo displaced
through the relief valve using the relation:

d: = 1~0 [tvo -(nr2 -f) I] % (4.18)

where V was the original specimen volume, f the compressibility


factor of the oil obtained from tables, r the radius of the ram, F the
axial force acting on the ram, E the modulus of elasticity of the ram
steel and I the measured axial displacement of the ram. Axial strain
was obtained from the axial displacement of the ram and the original
specimen length, and axial force from a load cell mounted above the
triaxial cell.
The cell was designed to test a 75 mm diameter by 150 mm long
specimen - an aspect ratio not, strictly speaking, large enough to
eliminate end effects during testing. Choice of specimen size and
aspect ratio is often an arbitrary process. The specifications tend to
'--I"""'" L J . _ TOP (OVER

SPECIMEN
SLEEVES

Figure 4.11 Section through a triaxial cell designed to accommodate large


axial strains and to measure volumetric strain using a bleed valve set to release
fluid at a constant confining pressure.
Rock Strength and Yield 97

recommend a minimum diameter of 54 mm and aspect ratio of 3.


Data in Fig. 1.1 on uniaxial specimens collected by Hudson et al.
(1971) illustrate some of the effects. Other features of the triaxial cell
are:

(a) a simple locating point and socket for easy central mounting of
the specimens;
(b) an unbalanced 250 mm long loading ram, allowing large axial
deformations and having a displacement of 45.6 cm 3 per cm of
travel;
(c) a radial margin of 22 mm;
(d) a threaded top cover, allowing quick and easy preparation ofthe
equipment for testing;
(e) a simple conversion for uniaxial experiments - by removing the
relief valve and pressure gauge from the system and replacing the
air bleed by a simple overflow pipe.
The cell mounted in the testing machine is illustrated in Fig. 4.12.

4.5 Axial and volumetric strain data

In order to illustrate some of the points made earlier, the results of


triaxial tests on seven rock types - predominantly weak rocks - are
reproduced in Figs 4.13 to 4.19. The results are described in Price
(1979) and Farmer (1980). Each of the figures includes stress-axial
strain curves to 10% axial strain or more and corresponding axial
strain-volumetric strain curves. Each figure also includes photo-
graphs of the deformed specimens after test.
Tests were carried out at some or all confining pressures between 0
and 42 MN m - 2 using the cell described in the previous section. The
value 42 MN m- 2 (approximately 6000 psi) represents the in situ
stress level at a depth of about 2000 m.
During the tests the cell was mounted in series with a load cell
(Fig.4.12) in a 5000 kN servo-controlled testing machine (see
Section 3.1), programmed to apply axial compression at a constant
strain rate of 21 J.1f. s- \ and capable of rapid unloading during
specimen collapse to allow study of post-peak strength charac-
teristics.
All specimens were prepared in accordance with ASTM Standard
02664, and each specimen was mounted between two platens for
testing and sheathed in a double flexible rubber membrane capable of
98 Engineering Behaviour of Rocks

Figure 4.12 Triaxial cell, load cell and bleed valve mounted in testing machine.

withstanding axial strains in excess of 20%, secured at each platen by


two rubber O-rings as in Fig. 4.11.
When testing, the triaxial cell was first purged of air and then
hydrostatically loaded to the required confining pressure by means of
an auxiliary hydraulic pump operated in step with the testing
machine. At the required confining pressure the relief valve was finely
adjusted and the confining system isolated from the pump by means
of a valve.
The rocks tested and the main features of their stress-strain
characteristics are described below. Geotechnical descriptions are
confined to unit weight, y, void ratio, e, and compressive strength, O'ef:
(a) Sandstone (y=25kNm-3, e=0.07, O'ef=97MNm- 2 )-a
medium-grained Coal Measures Sandstone. This was the
300

SANDSTO £
,,
N 200
i-

~v----
~

VI
~ ~21
go U)zlL

....
100
VI
'\ \. v---
~ I\.U3=4
I"-"~

...J

X
""'lC),7 v
~

t",-...
0
r-- "-U3-0

-
~
-.....: O"J 21

--
\ ~ ~ ..!Y]·r -m 14

""
~ -4
z 0)'4
:;;:
go
VI
.....
8 "- -..............
~>: ~p
::>
...J -..............
~
0
>

2 4 6 8 10
AXIAL STRAI N %

Figure 4.13 Axial stress-axial strain and volumetr;c strain-axial strain


curves for Coal Measures Sandstone specimens tested in triaxial compression at
confining pressures from 0 to 21 M N m - 2. 1n the photographs of fractured
specimens, test confining pressures were: bottom row, left to right, 4, 7,
7 MN m- 2 ,- top row, left to right, 14, 14,21 MN m- 2 (from Price 1979).
250
SILTY
SA DSiDN
200 LD
N
E
~ &~ ~.( ~ 2
~

ML~ ~
~ ISO

if ~
:!:
v'\
VI 0"3 _ 29 "v.,...
....~ lao
l
f-"
.....
J,
VI
0"3 :21

\ v 3 ·7

~ 2 ~
lT3.1;
6 0"3~O 8 10

- --
o
I~ 0"3_1
'"- 0"3-7 -<"0"3·36
<Yj:4
\.
"- ~ -
.......

2 ~
6
I---
6 6
r----. 10
AXIA STR AI %

Figure 4.14 Axial stress-axial strain and volumetric strain-axial strain


curves for Coal Measures silty sandstone specimens tested in triaxial
compression at confining pressures from 0 to 42 MN m- 2 • Test cotifining
pressures of the fractured specimens were: bottom row, left to right, 0, 4, 7,
14 MN m- 2 ; top row, left to right, 21, 29, 36, 42 MN m- 2 (from Price 1979).
200
(\

....
..!: 50
t\ VL
Y ,:
,, 'I
All VAtU S OFO"jA E1'1 MNAnZ

aJ. ze
"'"" ....,..,.......
::E:
VI
VI
... 100 I :, / '
I

g;
L ~ .~ ~ ~ , a) a 2

--..

L,
50 I
"""
-CT·lL'
~

<t
'""'--.
O).T
a),O
1'1', -28
.q3 -21
0].7".2- (Jl -14
z
« ~ r-..... PORTLA o STONE
~ -10
u
~
~
""~ -,5 ~
g r--- r---
-zc 6 8 10 12
AXIAL S AIN %

Figure 4.15 Axial stress-axial strain and volumetric strain-axial strain


curves for Portland Stone specimens tested in triaxial compression at confining
pressuresfrom 0 to 28 MN m -2. CO/ifzning pressures ofthefractured specimens
reproduced in the photograph were: bottom row, left to right, 0, 7, 7, 7, 14,
14 MN m- 2 ; top row, left to right, 14,21,21,28,28 MN m- 2 (from Price
1979 ).
;;:
700 1 ,. I ::>
>"-
<I
G) . 35
150 "";0;-
0:
N
E
GJ'18 ...
..... ° 3-11 '"'--'
L.
,.. 100 ~L
VI
'"....
0:
G3 · 1• -3
~ 0
VI >
..j
<{ SO ~= 7
X
<{

AL L VALUES OF O'~ ARE IN MN ml


t 1
I ' 1'8 ~B~O
oW"'I 2 L 6 6 10 6 S 10
AXIAL ST RAIN 'Yo AXiAl SfRAIN %
SACC~OAL LI MESTO NE SAC CfIAROIDAL LIMESTONE

Figure 4.16 Axial stress-axial strain and volumetric strain- axial strain curves for saccharoidallimestone specimens tested in triarial
compression at confining pressures from 0 to 42 MN m - 2 , In the photographs offractured and deformed specimens. test confining
pressures were: bottom row, left to right, 0, 7, 14 M N m- 2 ; top row, left to right, 21, 28, 42 M N m - 2 (from Price 1979),
104 Engineering Behaviour of Rocks

strongest rock tested. Difficulty was experienced in testing this


rock to large axial strains, owing to failure of both the inner and
outer rubber membrane sleeves in several tests. At low confining
pressures (up to 14 MN m - 2), predominantly vertical fractures
occurred at peak stress (Fig. 4.13) parallel to the minor principal
stress. At higher confining pressures, shear planes were formed.
The stress-strain curves demonstrate a rapid fall from a peak to a
residual stress level accompanied by rapid dilation at low
confining pressures and representing brittle or strain-softening
behaviour. The rapid and large dilation associated with peak
stress should be particularly noted.
(b) Silty sandstone (y =23.4 kN m -3, e=O.l1, O"cr=61 MN m -2) - a
medium-grained Coal Measures Sandstone with silt inclusions.
This deformed in a similar way to the sandstone, although shear
and conjugate shear planes (Fig. 4.14) were much more evident at
higher confining pressures. The stress-strain curves show a less
rapid fall from peak to residual strength and reduced dilation at
higher confining pressures.
(c) Portland Stone (y = 22.9 kN m- 2 , e=0.14, O"cr=73 MN m- 2 ) - a
strong, homogeneous spheroidal Jurassic Limestone. All speci-
mens exhibited strain-softening behaviour, i.e. the stress level of
the yield point decreased with increasing straining. Typically, the
stress dropped to some residual level at about one-half or two-
thirds of its peak stress in the case of the triaxially tested
specimens. All specimens failed in a brittle manner and most had
conjugate fault planes transecting them (Fig. 4.15). Difficulty was
again encountered with membrane failure at higher pressures.
(d) Saccharoidal limestone (y=26.7 kN m- 3, e=O.04, O"cr=49
MN m - 2) - a marmitized Carboniferous Limestone altered
by contact with a dolerite sill with a strong microstruc-
tural similarity to some classic marbles which exhibit ductile
characteristics, particularly that used by Scholz (1968a,b) and
discussed in the previous section. Its failure (Fig. 4.16) by tensile
fracture at zero confining pressure, through conjugate shear at
low confining pressures, by ductile faulting at higher confining
pressures and almost by homogeneous deformation at a confin-
ing pressure of 42 MN m - 2 illustrates the full range of rock
behaviour. This is mirrored by the stress-strain curves which
show a transition from strain-softening behaviour to near-strain-
hardening behaviour, although it should be noted that this is
accompanied by dilation.
150

N
MUDSTONE
0") • 0
~ -v ~
~
~

N
100
"}
I~
~).3S
l:
~. 28 """':"
Vl
fCr~'2
~

--
....
'" r~~ ~
...
a:
Vl 50
0"3 .1 4
f'-

--'
<t V- ....... 0").1
X
<t
CT) . O,

~
- <73 .1 (J).40

"'- ~
z
<i
~ 10
'"u
a:... 5
~ .O
....
l:
::::l
~
--'
~ 20 ~
'--.
2S 6 a
2 ~c
AX AL 5TRAI~ %

Figure 4.17 Axial stress-axial strain and volumetric strain-axial strain


curves for Coal Measures Mudstone specimens tested in triaxial compression at
corifining pressures from 0 to 40 M N m - 2. Confining pressures of the fractured
and deformed specimens reproduced in the photograph were : bottom row, left to
right, 7, 14, 21 M N m - 2 ; top row, left to right, 28, 35, 40 M N m - 2 (from Price
1979 ) .
0

/
"0') .1.2
MAR ..

160
(/
z
:>:: "Cr) =28
V>
40

(f::
Vl
C;! 0")-14

'" 0")- 7

20
'" ~
<l
>< G)"35
'" 1'-0"). 2 / ' Va=)· 0
o

~
1~
5 7
!,
) ·3-5 'lIT) < 7

"---
~
-5
~
-c
~XI A L '> T ~AI ~ %

Figure 4.18 Axial stress-axial strain and volumetric strain-axial strain


curves for marl specimens tested in triaxial compression at confining pressures
from 0 to 42 MN m- 2 . In the photograph of deformed specimens, confining
pressures were.' bottom row, left to right, 0, 3.5, 7 M N m - 2; top row, left to
right, 14,29,42 MN m- 2 (from Price 1979).
AXIAL STRAIN %

Figure 4.19 Axial stress-axial strain and volumetric strain-axial strain


curves for rock salt specimens tested in triaxial compression at confining
pressuresfrom 0 to 42 MN m- 2 • Deformed specimens were tested at confining
pressures of: bottom row, left to right, 0, 3.5,7,14 MN m- 2 ; top row, left to
right, 21, 28,35,42 MN m- 2 (from Price 1979).
108 Engineering Behaviour of Rocks

(e) Mudstone (')'=26.3 kN m -3, e=0.02, O"cf= 55 MN m- 2 ) - a fine-


grained Coal Measures Mudstone with occasional bands of silt
and ironstone. Deformational behaviour was similar to the
saccharoidallimestone, although the ductile behaviour is not as
clearly demonstrated (Fig. 4.17), partly because of the layered
structure of the specimens.
(f) Carnallite marl (,),=22.4 kN m- 3 , e=0.05, O"cf= 10
MN m - 2) - an anhydritic marl over Permo-Triassic evapor-
ites. Deformational behaviour was characterized by a pro-
gressive change from 'overall' strain-softening to strain-
hardening with increasing confining pressure, and from con-
jugate shear of the specimen to uniform flow (Fig. 4.18).
Although, as in the case of mudstone, the layered structure
inhibited homogeneous deformation, evidence of ductile
behaviour can be seen.
(g) Rock salt (')'=21.7 kN m- 2 , e=0.05, O"cf=26 MN m- 2 ) - a
Permo-Triassic rock salt with some minor marl layers. This
behaved in a ductile manner in all but the uniaxial experiment
(Fig. 4.19). It can be seen that in all except the lowest confining
pressure experiments, the specimens deformed homogeneously.
Deformation was however accompanied throughout by dilation.

It is possible to present the data in various ways. Initially, starting


from the discussion of yielding in Section 4.2, the deviator stress is
plotted against the confining pressure in Fig. 4.20 and compared with
Barton's (1976a) suggested critical state lines. The data plotted are
based on peak stresses and, in the case of rock salt, stresses at 2 %
strain. It can be seen that, of the rocks tested, only the mudstone,
carnallite marl and rock salt approach the critical state line, and only
the rock salt and carnallite marl reach the constant deviatoric
stress/confining stress state which defines the Tresca yield condition.
This approach is, however, too simplistic and a better indication of
the deformation characteristics can be obtained from Fig. 4.21 where
the peak and residual strength envelopes for six of the rocks (the
strongest, sandstone, is excluded) are plotted in q, p space. The
stronger rocks have distinct peak and residual envelopes, with an
indication that these will join at a higher p. In the weaker rocks this
joining takes place to be suceeded by a single line representing the q, p
relation during strain-hardening deformation. But this strain-
hardening deformation is still dilatant. As in the case of Scholz's data
in Fig. 4.9, if the data for rock salt are plotted in q, p space for different
Rock Strength and Yield 109

PORTLA 0 STO E
ROCK SALT·'lII2'10 £,
Mue STONE
~ SACCHAROIOAL LIMES 0 E
5 SILTY SA OSTO E
6 SA OSTO E
7 MARL

Figure 4.20 Differential stress against confining pressure for the rocks in
Figs 4.13-4.19.

dilations, it is again possible to obtain a family of yield curves


(Fig.4.22) representing different magnitudes of dilation for the
combined residual/peak curve. This is not the case with post-peak
dilation for the strain-softening rocks. Fig. 4.23 shows that for
Portland Stone all dilations follow a single linear relation represent-
ing the residual strength envelope of the rock.
There is therefore a case for describing the joining of the peak and
residual curves as the brittle-ductile transition or stability line to
denote the point at which deformation becomes stable. As explained
previously this is the point where energy input is required to induce
further deformation. This can form the basis for a simple, general yield
q .~ ./
150 ./' q
150
PEA~
.
.
100 100

50
/ / RESIOUAl

50
./ .
PORTLAND SILTY
STONE SANDSTONE
./

y;
15 'i() 15 100 P 15 'i() 100 P

~;/1
100 .1)-
125

100
./ 15

...
,j
15 / RES10UAL RESIDUAL
50

50
SACCHARQIOAl UOSTONE
LIMESTONE IS

25 'i() 75 oP 15 SO 1S P

31

15

CARNALliTE
MARL ROC~ SALT

15 50 P 25 50 p

POSI • PEAK
q. P CURV!S AI I%SIRAI q • 0, ·0, MNI Z
p. 'I, I C,
. 20,1 MNl m'

Figure 4.21 Peak and residual strength envelopes in q, p spaceJor the rocks in
Figs 4.14-4.19.
Rock Strength and Yield 111

100
q
MN/m 2

6V
v

100 P
Nftn2

Figure 4.22 Constant volumetric strain yield paths Jar rock salt calculated
Jrom data in Fig. 4.19 (Jrom Price and Farmer 1980).

criterion for rocks outlined in Fig. 4.24. This illustrates the changing
deformation characteristics of rocks having a peak strength envelope
describing the intact strength of rocks at low confining pressures. The
residual strength envelope will describe the shear resistance of either
(loose) broken rock or rockjill or the resistance of planar discon-
tinuities. The stability line marks the onset of ductile behaviour and the
critical state, the state of plastic yield. The general criterion is useful in
that it isolates the H vorslev surface, a series of yield curves of different
dilation describing the part ductile, part non-ductile deformation of
weaker rocks or strong rocks at high confining pressures. It is
sufficiently important in rock mechanics to warrant further exam-
ination.

4.6 The Hvorslev surface in rocks

A test of plastic deformation is whether the deformation of a material


can be specified by a yield surface, a flow rule and a hardening law.
100

SYMBOL AV'V
0-001
000.
0010

0411
~l

O~N

0~lO

CI'. '11 10 , • 20ll


I'1H I.l

Figure 4.23 Different volumetric strains plotted in q, p space for residual


shearing of Portland Stone calculated from data in Fig. 4.15 (from Price and
Farmer 1979).

CRI TICAL STATE LINE

RESIDUAL STRENGTH SURFACE

Figure 4.24 A simple representation of a general strength and yield criterion


for rocks (after Price and Farmer 1980).
Rock Strength and Yield 113

The yield surface defines the complete yield regime for a material
from the yield curve, representing the states of strps: ?t which yield
first occurs, to the failure envelope. In other words it separates the
states of stress causing only elastic deformation from those causing
both elastic and plastic deformation. The flow rule relates the ratio of
the increments of plastic strain during deformation - in other words
the plastic strain increment vector - to the states of stress causing
deformation. The hardening law relates the stress increment causing
deformation to the magnitude of plastic strain.
When a material flows plastically, it is possible to define lines of
equal 'velocity of flow' and equal 'plastic potential' through a plastic
potential function. Since the material will tend to flow in the direction
of the stresses causing yield, the plastic potential function must be
related to the state of stress. In plastic theory therefore, it is
conventional to assume that the plastic potential function and the
yield function defining the yield surface coincide and that the
deformation vector, normal to the plastic potential function, is also

DfFORMATIO VECTOR

~YIELO CURVE g
PLASTIC POTE TIAl'
FUNCTION

p
Figure 4.25 Representation oj the normality condition in q, p space.
114 Engineering Behaviour of Rocks

normal to the yield surface. This condition is known as the normality


condition and the flow rule relating ~tress to deformation at this
condition is known as the associated flow rule.
In Fig. 4.25 the normality condition is illustrated for a yield curve
plotted on q, p axes and it can be seen that in this case the normality
condition can be expressed as:
dq bv/v
(4.19)
dp bss
and the equivalent flow rule (see Atkinson and Bransby 1978) can be
written as:
q' =M _ bey (4.20)
p' bes
where M is a frictional constant and Bv and Bs are respectively
volumetric and shear strains, having different signs.
This is essentially the stress dilatancy equation for soils of Taylor
(1948) and Rowe (1962), and the basic form of the Hvorslev surface. It
can be demonstrated that rocks deforming in a stable manner obey
this equation by plotting, in Fig. 4.26, the ratio q/p against - bev/bss
for the experimental data for rock salt (Fig. 4.19) and carnallite marl
(Fig. 4.18). In addition Scholz's (1968a) data (Fig. 4.5) and data
obtained by Edmond and Patterson (1972) on Carrara Marble are
included. The data are summarized in a general form in Fig. 4.27.
The following points can be observed generally from Fig. 4.26 and
by comparing the data plotted in Fig. 4.26 with equation (4.20).
(a) There is in each case a linear part of the curve indicating a
H vorslev surface between the brittle-ductile transition and the
critical state (Fig. 4.27). Some recent experimental work by
Gowd and Rummel (1980) on a porous sandstone has shown that
the brittle--ductile transition and the critical state appear to be
coincident and by implication the Hvorslev surface, as defined by
equation (4.20), does not exist. Such behaviour appears to be an
exception to the general findings and is probably due to the
porous structure of the rock.
(b) Given that q = 0" 1 - 0" 3 and p =!(O" 1 + 20" 3), the maximum value of
q/p is 3 corresponding to the uniaxial state (0"3 =0), and forms an
upper bound for the curves (Fig. 4.27).
(c) The intercept of the curves in Fig. 4.26 with the ordinate gives the
value of the frictional constant, M. In these examples, ap-
proximate values of M range from 0.3 to 0.9.
Rock Strength and Yield 115

(d) The transition from stable to unstable deformation is marked by


a distinct deviation of the curve from linear behaviour; in other
words, equation (4.20) is only valid for stable deformation.
(e) The approximate slope ofthe linear part ofthe curves in Fig. 4.26
is respectively (a) 8, (b) 6.2, (c) 0.3, (d) 1.7, instead of unity as
implied by equation (4.20).
(f) The ordinate intercept &V/&5 =0 marks the boundary between
the H vorslev and Roscoe surfaces, i.e. the critical state, where there
is no further change in volume.
(g) The existence of a Roscoe surface is suggested by the curves in
Fig. 4.26(c) and (d) particularly at the very high confining
pressures used in these experiments.
(h) The very large scatter of data, particularly in Fig. 4.26(c) and (d),
indicates the difficulty of measuring small volumetric strains at
high confining pressures.
The form of equation (4.20) assumes that all the work transferred
per unit volume of the specimen during plastic yiel~ng is expressed as
friction. Gerogiannopoulos and Brown (1978) prbposed an alterna-
tive form to allow for the existence of other yield and non-yielding
mechanisms in rock deformation:

(4.21 )

where K is a factor describing other rock deformation processes.


In the post-yield condition which is considered in Fig. 4.26, other
factors affecting yielding will tend to require more energy than that
needed to induce purely frictional flow, in which case K will be greater
than unity. Before true frictional flow can occur in rocks, complete
cataclasis in a homogeneously deformed specimen must occur. It is
significant that Scholz (1968a) describes this type of breakdown in his
experiments on marble where K is closest to unity (Fig. 4.26(d)). In
the case of rock salt (Fig. 4.26(a)) and carnallite marl (Fig. 4.26(b)) it
can be seen (Figs 4.18 and 4.19)that cataclasis plays only a minor part
in the deformation process. In the case of the carnallite marl it is likely
that the high value of K results from movement along contiguous
shear planes. In rock salt the deformation process is more complex
with a possibility that part of the deformation results from irregular
intracrystalline movements which do not result in structural break-
down. It is significant that rock salt is one of the rocks to which
engineers would ascribe the property of creep or time dependency.
P (a) ROCK SALT

-... --
3
...
... ... ..-
/ "f M

·
/
(
1 .. + 3-5
J
J •
1 0 21
8
10
,0
• 42

1 x
x
J ~
I x

I
I
0 0-5 1 1-5 2

(c) CARRARA MARBLE


,-

-. . .
2 .,-

.,- ~
./
, .,
-s • to- ~~ ".

.
/.iJ ..
.... . . .....
.
+
M

x "' x
.... x
)It.
~ . ·,· 50
100
...." -# 200

7Jfl • 400

·
x 600
6' BOO
flt.
00
O-S

0 I 2 3 4 5
q/p j ( b ) (ARNALL ITE MARL

,
3
/

I
I
I M
I
·
..·
2 I
I
I
I
.. ,..Ii:
t
x

I'
I
1
I
x ~
I
I

0·5 ' ·5 6 Ev
- 6E s

p ( d) MARBLE
3

.- -
--- -...-
-

-
,
M
.,
2

/
/ ·,· o25
/ 50
~ + • 7S
"" /.1: .• +

·
+ A 100
1
)(1(~ X 200

)(,

, " ¥-
x
iX ' A

/ ~.
/0
,
0
0
A
1 2 3 4
-

Figure 4.26 The H vorslev surface in q/ p, fJEv/ &. space for (a) rock salt, (b)
carnallite marl, (c) Carrara Marble, and (d) marble (from Price and Farmer
1981 ).
118 Engineering Behaviour of Rocks

UNIAXIAL STATE
- - - - - - - - - - - - --...-- -- - - - - -
UNSTABLE
[)[FffiMATIO
_ _ _ _ _ _ _ _ _...._ _ _ SRI TTLE DUCTILE TRANSI TlON
OR STABIlITY LI NE
STABLE
[)[FOR I1ATlON

HV LEV
SURFACE

I • CRITICAL STATE. q/p = 11 . t. (. = 0


ROSCOE
SURFACE

Figure 4.27 A generalform of the Hvorslev surface (after Price and Farmer
1981 ).

This is a concept which has deliberately been ignored up to this stage,


except insofar as strain rates have been specified for tests and their
effect on strength mentioned. In soil mechanics time dependency,
except insofar as it is related to consolidation processes, is ignored. In
some rocks it may be overemphasized, but it is certainly an important
aspect of laboratory testing.
5 Time
Dependency

Time-dependent deformation in soil mechanics is expressed mainly in


terms of hydrodynamic time lag, or consolidation. The concept is
described elegantly and intuitively by Lambe and Whitman (1979)
who say that the time, t, required for consolidation should be:
(a) directly proportional to the volume of water squeezed out ofthe
soil, which will in turn be directly proportional to the product of
the change in confining pressure (~a), the compressibility of the
soil matrix (C) and the thickness, of the soil mass (H) equivalent
in one-dimensional terms to the volume;
(b) inversely proportional to the rate at which the water can flow
through the soil, which will be given by Darcy's law
v=ki=k ~a/H where k is the permeability coefficient and i the
hydraulic gradient.
Thus a relation can be obtained for t in the form:

~aCH CH 2
tock~a/H =-k- (5.1)

which is the same as Terzaghi's equation for 90% complete one-


dimensional consolidation - derived in all soil mechanics textbooks.
It is assumed that, apart from some secondary consolidation, all
deformation with the exception of the hydrodynamic time lag takes
place immediately upon application of stress. This is also the
assumption in rock mechanics, and yet there are indications that
there are time-dependent effects in rock testing and this is why a
standard test rate is specified. In weaker rocks, particularly, even

119
120 Engineering Behaviour of Rocks

when they are dry, there are indications of some time-dependent


effects which are worth investigation.

5.1 Creep strain

Hardy et al. (1969) have shown that if a specimen in uniaxial


compression is subjected to a constant maintained load in the
microfracturing range (see Fig. 3.15) the specimen will continue to
deform after initial application of the load. In addition there will be
continued microseismic activity (see Fig. 3.8) and the cumulative
number of events will be related to the strain of the specimen. This
type of strain is known as creep strain, and the relation between creep
strain, microseismic events and the level of applied stress obtained by
Hardy et al. (1969) for sandstone and limestone specimens is
illustrated in Figs 5.1 and 5.2.
The experimental technique in these tests was slightly unusual in
that the axial stress on the unconfined specimen was increased
incrementally by about 10% of the strength level every 40 min. It is
probable, however, that the data are representative of what might
normally be expected to happen during maintained load testing.
In the case of the sandstone, which has a compressive strength of
around 150 MN m- 2 there are several similarities between 'conven-
tional' brittle behaviour during strain-controlled loading, and
maintained loading. For instance, at the higher maintained stress,
increases in axial stress with time are accompanied by strongly
dilatant behaviour and by increases in microseismic activity which
are related to axial strain. It should be noted that axial strain is
increasing linearly with time at these stresses and it is likely that if the
test had continued fracture would have occurred.
In the case of the limestone, which had a uniaxial compressive
strength of about 50 MN m - 2, the deformation features, although
present, are less pronounced. There is nevertheless a tendency
towards linearity in the stress-axial strain curve and a linear relation
between accumulated activity and axial strain.
The relevance of this type of testing in rock mechanics will be
discussed later, but its most common use and that for which it is most
easily justified is in estimating the performance of pillars - often at
depth - in weak rocks, especially evaporites. Fig. 5.3 illustrates
typical creep curves for tests on rock salt specimens with a high
width/height ratio. The time-dependent strain is much higher than in
( 9S % COMPRESSIVE STRENGTH

Ire 5.1 Typical plots ofaxial and volumetric creep strain and accumulated microseismic activity against time, together with a plot of
Imulated microseismic activity against axial strain, for cylindrical specimens of Crab Orchard Sandstone loaded uniaxially in
pression at maintained stresses estimated as a percentage of uniaxial compressive strength (after Hardy et al. 1969).
\00 , , 10
<:
<l
~ e()
rOJ ...
- 6()
B
B
...
!1 '0
A 3 . ~
=>
:0::
=== ~ §
----
«

12 16 1() 21. 28 32 12 16 20 21.


TIME· MINUTES TIME: - MINUTES 28 32

100
""
..
:::
~
..
~
VI

V"I B
ffi A
~
a
eo
I~j
«
o
° 10
//
20 30 1.0 SO
TI ME - MINUTES AXIAL S 60 70
TRAIN - MICROSTRAIN

A _ 78% COMPRESSIVE STRENGT


e - 87% COMPRESSIVE STRENGT
C - 95 % COM PRESSIVE STRENG

Figure 5.2 Typical plots ofaxial and transverse creep strain and accumulated microseismic activity against time, together with a plot of
accumulated microseismic activity against axial strain,for cylindrical specimens of Indiana Limestone loaded uniaxially in compression
at maintained stresses estimated as a percentage of uniaxial compressive strength (after Hardy et al. 1969).
STRESS 21 MN/m2

2S

;::15
'-'
""
a:
~

g
VI

--'
<t
X
<t
OS

r-
o 10 20 3D 40 so 60 10 &0 90 100
nHEIOAYSI

Ire 5.3 Plots of axial strain against time for specimens of rock salt having a diameter of 150 mm and a width/height ratio of 2.
124 Engineering Behaviour of Rocks

the case or the stronger rocks tested by Hardy but the characteristic
shape of the curve is similar. In this case it describes either
logarithmically decaying strain with time or steady-state strain
showing an increasing linear relation between strain and time.
These typical types of deformation can be used to propose three
types of characteristic strain-time curve under maintained stress,
which can in turn be related to the various stages of the unconfined
strain-controlled loading deformation process in Fig. 3.15. These are
designated (a), (b) and (c)in Fig. 5.4. The basic mechanics of each may
be described as follows:
(a) If the level of maintained stress is above the critical crack density
level, rapid spreading of ,unstable' fractures will lead to accelerat-
ing creep strain and rapid fracture of the specimen.
(b) If the level of stress is well below the critical crack density level,
then there will be minor spreading, probably at an exponentially
decaying rate, of 'stable' micro fractures which will cause very
small creep strain with no failure.
(c) There may be an intermediate zone, just below the critical crack
density level, where continued spreading of 'stable' microfrac-
tures, albeit at a decaying rate, will approach the critical crack
density level and 'cross over' leading to 'unstable' crack
propagation and accelerating creep and failure.

£.

Figure 5.4 Postulated strain-time curves at (a) very high maintained stress
levels, (b) moderate maintained stress levels and (c) high maintained stress
levels.

The latter type of deformation would describe the commonly quoted


creep curve for rocks in compression comprising:
Primary creep - at an exponentially decaying rate:
eoclog t
Time Dependency 125

Secondary creep ~ or steady-state creep:


soc t

Tertiary creep ~ at an accelerating rate:

The level of strain above which creep may lead to failure varies with
different types of materials and with degrees of confinement, moisture
content and temperature. For stronger rocks, an ultimate long-term
uniaxial strength in the region of 70% of ultimate compressive
strength is often quoted (see Ladanyi 1974) and in weaker rocks with
high moisture content this can be reduced to 50--60%.

5.2 Phenomenological models of creep

There is a long history of maintained load testing in order to


investigate creep strain. The basic methods were described by Griggs
(1939) and refinements are discussed in detail by Dreyer (1973) and
Lama and Vutukuri (1978) among others. They usually involve a
simple test frame in which triaxially confined or unconfined
cylindrical specimens can be subjected to an axial compression force
from a fulcrum or oil reservoir topped up by a pump or by
compressed air cylinders.
In characterizing the time-dependent characteristics of rocks, there
are two common practices. The first is to characterize the strain in the
form of a phenomenological rheological model. These are essentially
viscoelastic models made up from basic deformation elements of a
spring (elastic deformation ~ Hookean material) and a dashpot
(viscous deformation ~ Newtonian substance) or viscoplastic models,
which include a frictional contact (plastic deformation above a yield
point ~ a St Venant substance). Examples of some of these models are
given in Fig. 5.5. By varying the number and configuration of the
elements, models of increasing complexity and accuracy may be
produced, with values of E, 11 and (Jy chosen to make the model fit the
creep curve of the material. Examples of equations which may be
quoted relating creep strain, s, to maintained stress, (J, and time, t,
include:

Maxwell model,
(J (Jt
s=-+~ (5.2)
E 11
126 Engineering Behaviour of Rocks

i
101 LI EAR MODELS I bl NO ·lINEAR MODELS

~
11AXWELL BINGHAM
• S(o· elas h ( plash(
flUid so\Jd

KELVI N·VOIGT
visco-elastic
sol .d
lli:::"'Oec
n
~ sol.d
ZENER PRAGER
flu.d

~
'¥"""
.. BURGHERS

Figure 5.5 Examples of rheological models comprising elements of H ookean,


Newtonian and St Venant substances which can be used to simulate the
phenomenon of creep under maintained load in uniaxial compression.

Kelvin-Voigt model,

(5.3)

Zener model,

a [ 1- E2 exp ( - -E1E2t)]
e=- --- (5.4)
El El +E2 I](El +E2)
where El refers to the solitary spring
Burgers model,

e=~ +~
E2 El
[1 - exp( _ E 1
1]1
t)] + 1]2at (5.5)

where E l' I] 1 refer to the spring and dash pot in parallel


Bingham model,
a
E
e= { (5.6)
t a
(a-a )-+-
Y I] E
Time Dependency 127

A second approach which is essentially the same as the first is to fit


some form of strain-hardening, time-hardening or ageing curve to the
data obtained from maintained load tests. These are based on a best-
fit basis and are essentially empirical. They have - together, to a lesser
extent, with the rheological models - the basic disadvantage that
since no fundamental mechanical or physical description of rock
deformation is present, a confusing number of relations can be
proposed for a similar curve. Mirza (1978) in a light-hearted but
necessary exercise suggested 24 equations which could be made to fit
the type of data in Fig. 5.3. These are reproduced in Table 5.1.
eruden (1971) also highlighted the problem and suggested the
rejection of the exponential equations to fit transient creep curves,
replacing them with a structural description of stress-aided corrosion
at the tips of microcracks. The implication is that any creep law
should be based on the physics or mechanics of deformation. This led
to some experiments on rock salt in the author's laboratory (Price
and Farmer 1981; Gilbert and Farmer 1981; Farmer and Gilbert

Table 5.1 Some common creep laws (after Mirza 1978)

1 e=Atrn
2 e=A +Btrn
3 e=A+Bt+Ct"
4 e=A+B~+C~+D~
5 e=Atrn+Btn+CtP+Dtq +···
6 e=A log t
7 e=A+B log t
8 e=A log(B+t)
9 e=A log(B+Ct)
10 e=A +B log(C+r)
11 e=A+Blog(t+Dt)
12 e=At/(1 +Bt)
13 e=A+Bsinh(Ct")
14 e=A+Bt-Cexp(-Dt)
15 e=At+B[I-exp( -Ct)]
16 e=A[I-exp( -Bt)] +C[I-exp( -Dt)]
17 e = A + B log t +Ctn
18 e=A+Bt"+Ct
19 e=A+Blogt+Ct
20 e=log t+Bt"+Ct
21 e=A log[l + (t/B)]
22 e=A[I-exp(B-Ct")]
23 e=A[I-exp( -Bt)]
24 e=A exp(Bt)
128 Engineering Behaviour of Rocks

1981) aimed at providing an alternative basis for the description of


time-dependent deformation.

5.3 Time-dependent deformation

There are two alternative methods to creep for describing time-


dependent deformation:
(a) axial stress relaxation to maintain a constant specimen length
under constant confining pressure and temperature conditions;
(b) variation of constant strain rates during conventional triaxial
testing.
The former gives results which are essentially a 'mirror image' or
reversal of maintained load tests, with the added drawback that
changes in pressure are more difficult to measure than changes in
strain. Fine et al. (1979) and Peng and Podnieks (1972) give examples.
It is still relatively uncommon to describe time dependency on the
basis of constant strain tests although the strain rate sensitivity of
rock strength (Patterson 1978 quotes numerous examples) has long
been recognized through recommended standard strain rates for rock
testing. The effect of strain rates on rock strength can be illustrated by
considering the strain history of two specimens of a similar material in
Fig. 5.6. One is loaded rapidly, the other in stages with a time interval
between each stage. There will be little difference during the initial
stages up to about 50 %of ultimate compressive strength. Beyond this
point, however, there will be increasing divergence with increasing
stress as 'creep' under quasi-maintained load conditions similar to
those in Figs 5.1 and 5.2 increases with increasing load increments
during the time intervals. The effect should be to reduce the strength
and the secant modulus at failure and to increase the strain at failure.
Bieniawski (1970) has shown in a series of tests on sandstone

AXIAl
STRESS
a CRUPA MAINTAINED LOAD
SLOPE EO AT RAPID LOADI RATE
~CURV£D DEFORI'<ATlON CHARACTERISTICS
AT lOW RATE OF LeADING

(
AX IAL STRAI N

Figure 5.6 Postulated deformation characteristics at rapid and slow (or


incremental) loading rates.
Time Dependency 129

(Fig. 5.7) that these effects can be demonstrated in a quite strong rock.
In weaker rocks the effect can be demonstrated even more effectively.
Results of triaxial tests at confining pressures between 0 and
42 MN m - 2 on rock salt at a constant strain rate of 2 x 10 - 5 S - 1 are
illustrated in Fig. 4.19. In order to examine the degree of time
dependency through the effect of strain rate on the stress-strain curve,
these tests were repeated (Farmer and Gilbert 1981) on the same rock
salt at strain rates of 5 x 10- 3,2 x 10- 4 and 2 x 10- 7 S -1. The results
are summarized in Fig. 5.8 as plots of deviator stress against axial
strain.

'50 r-~--'--.---' OURATlO~


OF ES
"'e
i \O°r-7~~~~~~~
ia
~
VI

%STRA\N

N
E
Z
l:

b \20 ~oooo UJ
r
~ \\0 I~
Sc
30000
.......
&.-
'"t/i
x<f.
<t
\00

90
~

--
0'· 10'5 10'6 10'
f
r--.
' O'B 10'
20000
-=- :~OOO
10
Z
:::>
<!.!- SEC'
dt

;;:.
w

---
.... ,----
cr '·5
:::>
.... 1
~ f-"
~ 0·5
z
~ 0
....
V>
10" 10.5 10.6 10.7 10·a 10'9 10
RATE OF STRAI d tid! 5"

Figure 5.7 Injluence of strain rate on stress-strain curves, and uniaxial


strength, deformation modulus and strain at failure defined as peak stress, for
sandstone specimens tested in uniaxial compression (after Bieniawski 1970).
1)1) N 100
~
~ STRAIN RATE S. lO-ls ec - I z ,
Z l: 2x 10 - L 50'
l: 0): II> eo
80 (I) -
II> ~
~ --0:========== ~2 MNlm2 ...a:
a: II>
... ...,60 :1~ MNlm'
II>
U ~ : ii' ~1~152e
f?
0 ...;
...a< >
~
~
0
35 20. _
35

,
10 10
AX I AL STR AIN 1%) AX IAL STRAIN 1%)

100 50

N 2.10- 7 . . . -,
2 ,10 -5 sec -, 0):
"
z80
>- 315MNlm2
II>
(J)_ 21
II>
421
~ 60
l ~!
~ 30
==;;;;;;;~!!!!!,!!!!,,:~2e~221 14
VI ~ 1,35 MNlm' a: ~ -----====
..., Iii
0: ~ )5
u
~ a:
.. 0
~ :;:
c
~
0

, r -L--
8 10 0 2
AX IAL STRAIN 1%) AXIAL STRAIN 1% I

Figure 5.8 Deviatoric stress-axial strain curves for rock salt specimens tested in triaxial compression at confining pressures between 0
and 42 M N m - 2 and strain rates between 5 x 10 - 3 and 2 x 10 - 7 S - 1 (after Farmer and Gilbert 1981).
Time Dependency 131

(0)

(b)

Figure 5.9 Specimens of rock salt after testing in triaxial compression at strain
rates of (a) 5 x 10 - 3 S - 1 and (b) 2 x 10 - 5 S - 1. Confining pressures were.' 2R,
42MNm- 2 .. 2S, 35MNm- 2 .. 2T, 21 MNm- 2 .. 2V, 14MNm- 2 .. 2V,
7 MN m- 2 .. 2W, 3.5 MN m- 2 .. and R6, 42 MN m- 2 .. R5, 35 MN m- 2 .. R4,
28 MN m- 2 .. R3, 21 MN m- 2 .. R2, 14 MN m- 2 .. Rl, 7 MN m- 2 .. R8,
3.5 MN S-1 (after Farmer and Gilbert 1981).

It can be seen that at the two faster strain rates there is a tendency
towards strain softening at the lower confining pressures (3.5 and
7 MN m - 2). At the lower strain rates there is a tendency to strain
softening in the unconfined state only. Photographs of the specimens
tested at the fastest (5 x 10 - 3 S - 1 ) and slowest (2 x 10 - 5 S - 1 ) strain
rates shown in Fig. 5.9 illustrate the quite brittle nature of deform-
ation in the former case at the lower confining pressures.

5.4 Time-dependent strength reduction

In Chapter 4, the dangers of using the term 'strength' in connection


with rock deformation were emphasized. In Fig. 5.10 the stress
invariants taken from the data in Fig. 5.8 at axial strains of 2, 4 and
10% are plotted in order to allow comparison of 'strength' or yield
curves at various strain rates. Each of the curves follows the general
form of the envelope for rock salt in Fig. 4.21, having an origin in the
brittle field followed by an extended transition to ductile flow
132 Engineering Behaviour of Rocks

q M 1m'
70
q' 0 79p

60 60
2%
%
AXIAL
50 STRAI so AXI~l
STRAIN

40 40 ·

30 30
STRAIN RAn: (s .c'l
STI/AI RA T£ (so< - ~
y 5.'0- 3
20 51( 0-3
;II
o 2111wr4
C 2' la-I.
00 2KlO-s
o2.1Q-~
10 • 21( 10-7
. 2 .10- 7

1,0 60 80 1,0 60 80
P HIm' pMN /m'

q MN I 1

70

60
0°10
AXIAL
STRAIN
50

40

30
STRAI RATE I s.,-' I
, 5.10-3
C 2.,0-1.
00 2- O-s

20 1,0 60 80
P MIHII'

Figure 5.10 Strength/yield envelopes plotted in q, p space from the data in


Fig. 5.8 at 2, 4 and 10% axial strain (after Farmer and Gilbert 1981).

reaching a critical state at q/p values between 0.74 and 0.9. The effect
of the reduced strain rate is to enhance the ductile behaviour of the
specimen.
Gilbert (1981) has taken this a stage further by plotting the
intercept of the strength envelope and the critical state line in
Fig. 5.10, to demonstrate a linear relation in Fig. 5.11 between
Time Dependency 133

: BACK A ALYS IS
70 I
I
I
N
60 I
E I
I
z
E 50
VI
~
e:
VI
0

u
a: 30
8
«
~ 20
0

10_ 2
-3 -4 -5 -6 -7 -6 -9
LOG 101£ 1)

Figure 5.11 Plot of the deviatoric stress intercept with the critical state line in
Fig. 5.1 0 against the logarithm ofthe strain rate and supplemented by additional
data from back-analysis of variable strain rate tests described in Gilbert and
Farmer (1981) (after Gilbert 1981).

deviatoric stress and the logarithm ofthe strain rate. This can be used
to postulate, for deformation at the critical state, a unique relation
between q, p and log Il.
The implication of this is that the distinction made between the
short-term behaviour of rock salt expressed in terms of strength and
the long-term behaviour expressed in terms of time-dependent
deformation is to a certain extent artificial. The complete mechanical
characteristics of the rock would require measurement of volumetric
strain and would then appear as a series of yield surfaces similar to the
Hvorslev surface (see Fig. 4.26). The type of deformation is not of
paramount importance. Gilbert (1981) on the basis of the tests in
Fig. 5.8 has attempted to define the brittle and ductile fields for rock
salt in terms of confining pressure and rate of strain (Fig. 5.12) with
some success. The main point remains, however, that irrespective of
the exact form of any relation between q, p, dc/dt and dcjdc s ' rocks
such as rock salt which are recognizably time-dependent in their
reaction still retain sufficient strength or shear resistance (even in a
ductile state) to redistribute stress around an opening in a continuum.
The emphasis on time-dependent deformation in the phenomeno-
logical models developed for essentially 'soft', pillar loading con-
ditions can, when used in continuum mechanics, predict unrealistic-
ally high deformations. The emphasis in the mechanical descriptions
42 00 00 °0 KEY 00
o ~UCTILE

VC VERTI CAL CRACKS


'~ W WEDGE FORMATION
35 °0 °0
° DIVe- "--. TENSILE IBREAK~OWN
SP SH EAR PLANE FORMATION 00
OUCnLE IVFRTICAL ~
(RACKING
E 28 °0 °0
~
l:

.... ~,
'"=> DUCTILE
18c:: 21 OW~' 00
Q.
O OIV~
'"
~
WEDGE! DUCTILE
~ 14 . WIO ° OIVC 00
o ~, \ 00
'-'

~
• ______ ° W/ O ~ '\~D 00
WEDGE
----- - ~
OW ------. W ' ........... 0' SHEAR
__ -.... ........... velD °0 PLANE
o TENSION fRACTURE
t I - - - ~"'- -- - -->-;C~,- - - - . - - - i - - - - _ s pt?RMATlON
-3 -4 -5 -6 • I
-1
LOG 10 I e", I

Figure 5.12 Types ofdeformation observed during tests in Fig. 5.8 expressed in terms of confining pressure and logarithm of strain rate
(after Gilbert 1981).
Time Dependency 135

developed using strain-controlled tests similar to the 'stiff' loading


conditions which exist in a continuum is on stress redistribution
associated with 'strength' reduction as strain rates reduce.
Deformations and particularly lining stresses predicted using this
approach are significantly lower.
It is interesting to compare also the deformation in Figs 5.1, 5.2 and
5.3. All these deformation curves can be described phenomeno-
logically by the models in Fig. 5.5 and the equations (equations
(5.4)-(5.6)) describing them, or by the creep laws in T~bl~ ~,l! lind yet
it is quite clear that the deformation process in Figs 5.1 and 5.2 is
similar to, if not exactly the same as, the brittle deformation process in
uniaxial compression described in Chapter 3 and covered by the
'tension fracture' zone in Fig. 5.12. An increase in confining pressure
would modify it instantly. It is essentially a 'time lag' on the
deformation process, resulting in long-term strength reduction of the
type described by Ladanyi (1974). It is not ductile or viscoelastic
behaviour.
In Fig. 5.3 the confining effects of the higher width/height ratio will
undoubtedly have the effect of inducing ductility in the specimen. It
would however be simplistic to attempt to describe the deformation
of this sample solely in terms of time dependency. Fig. 5.13 illustrates
ductile and brittle behaviour in pillars of potash of different
width/height ratio subject to similar forces. That with the low
width/height ratio has deformed in a brittle manner. The other shows
signs of ductile behaviour at the centre, and brittle behaviour at the
outside. The resultant time-dependent deformation can be described
quite realistically in terms of stress redistribution in the pillar, leading
to a stable state and accompanied by various stages of brittle and
ductile deformation.

5.5 Cyclic loading

Another typical type of time-related loading which can occasionally


occur in rocks is cyclic loading. If a specimen is loaded and unloaded
there will be some degree of incomplete recovery. This will depend on
various physical factors and on the magnitude and rate of loading of
the rock. Cyclic loading in uniaxial compression should lead over one
or two cycles at (say) 50% of ultimate strength to an increase in
modulus and possibly a reduction in strength. At higher loads and
over a number of cycles it will lead to a significant reduction in
Figure 5.13 Types of deformation observed in potash pillars at a depth of
about 1100 m. Note that the pillar with the large width/height ratio (a) deforms
in a ductile manner. The pillar with a low width/height ratio (b) deforms in a
brittle manner.
Time Dependency 137

strength. This process, usually accompanied by some strain harden-


ing, is known as fatigue and the reduced strength level is known
as fatigue strength. It is analogous to the long-term strength ander
maintained loading discussed in the previous section, and there are
many predictable similarities between deformation under cyclic
loading and creep. The first of these is the relationship between the
ratio of fatigue strength and ultimate strength and the number of
cycles - known usually as the S-N curve (Fig. 5.14). Most rocks and
concrete have a fatigue strength between 60 and 70% of ultimate
strength.

ermax09
ere 0.8 \
07
--t
""'- I--.- r-::-. I=- FATiGUE
LI M!
06
10 10 2 10 3 10' 105
No OF CYE LES N

Figure 5.14 Typical fatigue or S-N curve for a rock or concrete material
relating fatigue strength, (im,x to uniaxial compressive strength, (ie/Jor a given
number of loading cycles.

The mechanics of deformation insofar as they may be postulated


are similar to those in creep, except that the cycling process represents
a direct energy input which would be expected to satisfy the
conditions for crack propagation (equation (3.4)) much more quickly
than under maintained load. The time duration to failure will depend,
therefore, on the amplitude and frequency of cycling.
This is illustrated by results reproduced from Attewell and Farmer
(1973) in Figs 5.15 and 5.16. In the first series of tests, 25 mm diameter
by 50 mm long dolomite specimens (Fig. 5.15) chosen for their
homogeneity were subject to cyclic loading at a frequency of 20 Hz
and at a controlled mean stress set at between 25 and 50 % of ultimate
strength and a stress amplitude maintained at a level sufficient to raise
the peak dynamic stress to between 40 and 70% of ultimate strength.
Maximum and minimum strains were recorded continuously during
loading. Shapes of typical curves can be compared with the shape of
creep curves (Figs 5.1-5.4).
In the second series of tests on a weaker dolomite sample,
specimens subjected to a mean stress of 50% of ultimate strength and
a stress amplitude to raise this to a maximum stress of75% of ultimate
1- 0 ~--~-~-~--------,

0·8 - - - - ! - --/-- - + - ----l

~ 0·6
.... !,-u
m,n
I I (Tm,n:335MN/m 2 )
""z I
~ I
t;; 0 I. ~----1-____!___I_--~------'
0-
t!l
""w
'-'

z~ 0 2 r-f+--t-r==::s~~~::;-;'j
>-
Cl

20 40 60
STRESS CYClESlt-;).10 - 3

Figure 5.15 Influence ofcyclic stress variation at constant frequency on strain


as afunction of time for dolomite specimens (after Attewell and Farmer 1973).
Note that initial elastic strains are discounted and that initial minimum and
maximum 'creep' strains are used as the zero base.

12---------~-~-~---~----------

1-0

0·8 100H z
~ 2·5Hz
~ 0·6
or 20 OH,
t;; 04 I N fo ,1, 1.1-0001
a.
w
.....
""
u 02

00 0·4 0·8 ' .2 1 .6 20 2.1. 28 )·2 3·6 40 I. -I. 4·8


STRESS CYCLES I '.10- 3

Figure 5.16 Influence of stress cycling frequency on dolomite specimens


subjected to the same mean stress and stress amplitude of cyclic loading (after
Attewell and Farmer 1973).
Time Dependency 139

strength cyclic loading was carried out at frequencies of 0.3, 2.5,10 and
20 Hz. The results (Fig. 5.16) show typical failure curves with failure
occurring in a lower number of cycles at the lower cycling frequencies,
but (within experimental error) in the same elapsed time of
approximately 1000 s, irrespective offrequency and at the same axial
strain.
Haimson and Kim (1971), in an interesting contribution, proposed
that the axial strain at fatigue failure (and by extrapolation failure
during maintained loading) is defined by the intersection of the upper
limit of cyclic stress (or the constant maintained stress) and the
unloading part of the complete stress-strain curve. The proposition
illustrated in Fig. 5.17 has a certain logic but must remain speculative
in the absence of more detailed information.

STRfSS
(J

FigureS.17 Relation, proposed by Haimson and Kim (1971), between a cyclic


loading regime and a strain-controlled stress-strain curve for rock.

5.6 Rapid loading

When a rock is loaded at a rapid rate, discussion of strength and


deformation must be approached from a slightly different standpoint.
Rinehart (1962) has shown that if a conventional cylindrical
specimen was subjected to a stress pulse of short duration, then the
tensile stress that it could withstand without fracturing was increased
by an order of magnitude. Others have shown that rapid application
offorces to rock leads to a relatively low transfer of useful energy. This
is particularly the case where explosives are used, when less than 1%
of the available energy is used in fracturing the rock. Much of the
study of dynamic forces and their effect on rocks is the province of
texts on rock blasting, stress wave propagation and rock crushing and
comminution. It is, however, useful to have a general understanding
of the mechanics of energy transfer from the application of rapid
loads.
140 Engineering Behaviour of Rocks

When an explosive is detonated in intimate contact with a rock


body, the instantaneous pressures generated can vary from ap-
proximately 50000 atmospheres to several million atmo5ph~r~5!
depending on the type and quantity of explosive involved and its
velocity of detonation. Some of the energy transmitted from the
explosive to the rock does work in pulverizing or melting the rock in
the immediate vicinity of the impact point whilst the remaining shock
energy passes directly into the rock body in an unstable compressive
shock front travelling at a speed greater than the sonic velocity of the
rock.
A true shock wave is only formed when the initial explosive pressure
so far exceeds the strength of the rock in compression that any plastic
state is bypassed and it can be said to behave hydrodynamically. In
other words, the pressure state in a rock simulates that in a liquid.
The shock wave has a very limited duration and rapidly attenuates
in both energy and velocity through a plastic phase to form a seismic
wave. This is an oscillatory compression pulse, propagating at a sonic
velocity and having insufficient energy permanently to disturb the
rock in its path. Seismic wave travel can therefore be analysed
theoretically by combining Newton's laws of motion with the theory
of elasticity assuming wave motion through an isotropic elastic solid
(see Jaeger and Cook 1969).
There are two basic types of seismic wave: body waves which travel
through the interior of the rock body, and surface waves which can
only travel along the surface of the material. Body waves can be
subdivided into two modes: compression or primary (P) waves and
shear or secondary (S) waves. Solving the general equations of motion
gives relations of the form:

Cp -
_(K +4/3G)1/2 (5.7)
P
and

Cs -_(G)1/2 (5.8)
P
for wave motion in three directions, where K and G are the bulk
modulus and the modulus of rigidity, p is the material density and C p
and C s are the wave or sonic velocities of the P- and S-waves.
These equations can be compared with the equations for strain in a
continuum (equations (2.33) and (2.34)) and can also be modified for
Time Dependency 141

the one-dimensional case:

(
E(I-v) )1/2 (5.9)
C p = p(l+v)(l-2v)

and

(
E )1/2
(5.10)
C = 2p(l+v)
s

or

(5.11 )

and
0.5 - (C S/C p )2
v=----;:- (5.12)
1- (C S/C p )2
These equations are the basic equations for computation of the
dynamic deformation modulus and Poisson's ratio from field or
laboratory measurements of P-wave and S-wave velocities, men-
tioned in Chapter 1, and various workers (cf. Judd and Huber 1962)
have demonstrated relations between sonic velocity in rock and the
square root of the deformation modulus. The P-wave and S-wave
velocities are affected by the structure and degree of saturation of
massive rocks, since air has a low sonic velocity. Typical P-wave
velocities are given in Table 5.2.

Table 5.2 Typical P-wave velocities

Cp p
Rock (m S-l) (Mg m- 3 )

Granite 3000-5000 2.65


Basalt 4500-6500 2.85
Dolerite 4500-6500 3.0
Gabbro 4500-6500 3.05
Sandstone 1400-4000 2.55
Shale 1400-3000 2.3
Limestone 2500-6000 2.5
Marble 3500-6000 2.65
Quartzite 5000-6500 2.65
Slate 3500-5500 2.65
142 Engineering Behaviour of Rocks

An oscillatory seismic wave in rock takes a form similar to that of


simple harmonic motion, the wave being defined in terms of
frequency, f, wavelength, A (where A= Cp/f), and the amplitude of
particle displacement, Ad. Ad can be related to the corresponding
amplitude of particle velocity, A v, through the frequency:
(5.13 )
Av is a useful indicator of rock movement in the wave path and can
be related directly to stress, strain and energy levels in the wave
motion:
(5.14 )

(5.15)

W= ae=pC~ (5.16)
The energy in the wavefront attenuates with distance from the
source. This is covered in detail in Attewell and Farmer (1976).
It should be stressed that, by definition, the relatively low-energy,
seismic compression waves will not damage rock through which they
pass although vibration of structures founded in rocks - particularly
at frequencies near to their natural frequency - can cause structural
damage. Reflection of seismic waves at surfaces - whether these are
present in underground structures or as quarry faces - which changes
the sign of the seismic wave from compression to tension, can cause
real damage. This can range from spalling from low-energy reflec-
tions, or in association with gas pressures released by detonation of
explosives, to shattering of the rock. This is the basis of blasting.
6 Discontinuities
in Rock Masses

Discontinuities play a major role in determining the behaviour of


rocks, particularly at or near exposures such as tunnel sidewalls,
slopes, open cuts and foundations. Where major, open discontinuities
exist, deformation may occur through movement along discon-
tinuities and through rotation of blocks rather than through
breakdown of the intact rock.
The term discontinuity is used widely in rock engineering to
describe any measurable interruption of a rock mass. It is often used
to the exclusion of geologically more acceptable terms such as
bedding, lamination, fault and joint, in order to emphasize the
importance of the existence of discontinuities in controlling the
engineering behaviour of rock masses, rather than their genesis. The
term fissure is sometimes used instead of discontinuity but this
normally has a specific use (Fookes 1965) to describe a small
terminated discontinuity. Systematic discontinuities are usually
referred to as joints. It is, in fact, difficult to imagine any series of
discontinuities as completely random. Terzaghi (1965) quotes
Bruno Sandor as saying that, on the basis of 50 years experience, ' ...
it is certainly easier to imagine random jointing for the purposes of
computation than it is to find an example of it in nature'.
Discontinuities appear therefore, invariably, in patterns or sets and
this is the reason for their importance in engineering design.
The recognition and analysis of discontinuities is an important
bridging link between the important concepts of rock material
behaviour discussed in previous chapters, and the actual performance
of rocks in the vicinity of structures in massive rock. It is, however,
important to recognize that, whilst it would be wrong to design such

143
144 Engineering Behaviour of Rocks

structures solely on the basis that rock is a material, it is equally


wrong to consider a rock mass merely as a series of discontinuous
structural units.
There are various ways of attempting to define the significance of
discontinuities, and three of these are illustrated in Figs 6.1--6.3. In
Fig. 6.1, Hoek and Brown (1980b) neatly illustrate the transition from
intact rock to heavily jointed rock mass with increasing sample size in
a rock mass surrounding an underground excavation. In Fig. 6.2,
John (1974) illustrates the essential difference between the major
discontinuities - probably sheared and continuous over a wide area,
and which are likely to cause major stability problems in a dam
foundation - and the minor structural and fabric elements. In
Fig. 6.3, Londe (1973) attempts to classify discontinuities according
to spacing and genesis and suggests a method of recording in the form
of histograms. Whilst useful in illustrating the significance of
discontinuities, they each indicate the nef!d for a rigorous method of
discontinuity description and measurement.

ROCK MASS

Figure 6.1 Il/ustration, after Hoek and Brown (1980b), of the transition from
intact rock to heavily jointed rock mass with increasing sample size around an
excavation.
Discontinuities in Rock Masses 145

ROCK ElEME TOR


MICRO HE ENT
STRUCTURE
MI OR JOIN NG

ROCK SUB SYSTEM OR


ACRO ELEMENT "..~=_ MAJOR GEOLOGIC STRUCTURE
co TAIN ING INOR Ie =CON TI NUOUS
Pl ANES OR
DISCONTINUOUS JOINTS ZONES OF WEAKNESS

Figure 6.2 Illustration, after John (1974), of the different effects of minor
fabric elements and major discontinuities on the potential stability of a dam
abutment.

SPACING
l km
:E
'-' Vl
u::
z
Q
w
J:
Vl
g
1
.•
100 m
10m
1m ...
~

I~
10em
1=r.r~:E
1em
tt.:' IQ'=
.q;w m
'-'Vl - '
>-

--........-:'4:
~
lmm t!t:i i=~
:::;t!= ~
O.1 Vl>-
z_ 0
a:
0·0 I ....
t-«
a:
0·00 1 0
co
~

Figure 6.3 Classification system for discontinuities suggested by Londe


(1973), together with suggested methods for determining rock behaviour.

6.1 Discontinuity measurement

Piteau (1970) has suggested seven similar discontinuity character-


istics having particular significance in engineering, and requiring
quantitative evaluation, which should form the basis of any
programme of discontinuity measurement. They are:
146 Engineering Behaviour of Rocks

(a) type of structure;


(b) position in space;
(c) orientation;
(d) intensity;
(e) surface hardness and roughness;
(f) continuity;
(g) type and thickness of any infill.

Collection of this information requires a systematic discontinuity


survey technique.
Robertson (1970) and Piteau (1970) were among the first to
describe discontinuity survey techniques. The purpose of a discon-
tinuity survey is to produce a reliable model of the discontinuity
patterns in a rock mass. Since it is often physically impossible to
examine every discontinuity in a rock mass, the information is usually
obtained from a sample of discontinuities. These samples must be
sufficiently large to construct an accurate model and, where
necessary, to detect regional variations in the discontinuity pattern.
The accuracy of sampling is considered later.
Discontinuity surveys can be carried out on various types of
natural and artificial exposure. In order of increasing area of exposure
sampled, these include borehole cores, large borehole sidewalls,
shafts, tunnels, trenches, artificial cut or slope faces and natural rock
outcrops. Each has disadvantages - in cores, orientations are not
usually possible; borehole sidewalls have limited areas; in artificial
excavations, it is sometimes difficult to differentiate between natural
fractures and fractures induced by excavation; and at outcrops,
discontinuities may be blinded and altered by weathering. Ideally,
surveys should be carried out in cuts, trenches or shafts, carefully
constructed in the strata where information on the rock mass
properties is required, so that access can be obtained for a careful
examination of a fairly extensive zone of freshly exposed rock.
Unless the exposure is severely limited, systematic sampling
techniques based on a limited areal or limited linear extent can be
used. Of these, the linear method of scan-line sampling is preferred
because the mathematics of data analysis is simpler and the method is
more flexible and economical. A detailed outline of the method is
given in Attewell and Farmer (1976), and of the field procedure in
Hoek and Bray (1974).
Ideally, the method requires three orthogonal scan-lines of equal
length and at least 50 times (see Priest and Hudson 1976) the mean
Discontinuities in Rock Masses 147

discontinuity spacing in length. The scan-line will normally take the


form of a steel tape pinned to the face. Then for every discontinuity
which intercepts the tape the following information will be recorded:
(R) digtance from the end of the tape;
(b) direction of true dip, 0-360° azimuth bearing;
(c) dip from horizontal;
(d) if the discontinuity terminates, the length of the discontinuity;
(e) width of the discontinuity;
(f) discontinuity surface and infill properties;
(g) shape of discontinuity and any other relevant features.
In a densely fissured face it may be necessary to select discon-
tinuities or to concentrate on size, and dip and dip direction data. This
can be justified because the resistance to movement along a potential
failure plane in a rock mass is determined by the respective
orientations and inclinations of the plane and any discontinuities in
the rock mass. Thus stability will be determined by the size, number,
orientation and spatial density distribution of discontinuities in the
rock mass.
Robertson (1970) and Terzaghi (1965) have considered some of the
sources of error in discontinuity surveys. These are unavoidable in
any collection of observational data and take two major forms: errors
of measurement and errors of selection. Errors of measurement
include misreading of dip directions at high and low discontinuity
inclinations. Robertson quotes a resurvey which indicated that
maximum errors of 5° in dip angle and 10° in dip direction could
reasonably be expected in a discontinuity survey. In the case of
surface and infill properties, where speed may require a classification
system, there may be large errors but these data are less important in
preliminary stability analysis. Errors of selection include missing of
small discontinuities; measurement oflarge discontinuities more than
once; overlooking of discontinuities at a low angle to the scan-line
and of discontinuities parallel to the bedding; and variation of details
of sample selection between different observers.
Corrections which can be made to standardize data are outlined in
Attewell and Farmer (1976). Particular corrections can be made for
scan-lines of unequal length, for scan-lines where a low number of
discontinuities are encountered due to weathering and for discon-
tinuities nearly parallel to the scan-line .
.R0be~tson (1970) also considers the accuracy of discontinuity
onentatIOn and spacing density estimation from a given survey
148 Engineering Behaviour of Rocks

sample. Since stability along a particular plane in a rock mass is


determined by the size and spatial arrangement of discontinuities
which occur along the plane it is important to estimate the accuracy
with which discontinuity densities can be determined. Robertson
shows by statistical analysis that about 100 observations of
discontinuities from a set are required to estimate the population
density of the set with 95% confidence to within ±20% oftrue value.
The basic conclusion of Robertson is that interpretation of
geological factors can yield sufficient information for preliminary
stability analysis of massive rock structures, but that the information
is relatively inaccurate. In particular, the data are weak where
information on the resistance to shear along a joint set is required
from data onjoint continuity, joint surface properties and intact rock
properties.

6.2 Discontinuity orientation data

The most important data defining the difference between rock masses
and intact rock are the spatial orientations of discontinuities. Unless a
potential failure plane or combination of failure planes approaches or
coincides with a plane or combination of planes containing a joint or
discontinuity set, then there is no obvious difference between a rock
mass and a continuum based on intact rock. The most important
approach to analysis must, therefore, be to determine the orientations
of discontinuity sets. This can be done subjectively by the geologist, or
by plotting data from discontinuity surveys on some form of
stereographic projection.
The development and use of stereo graphic projections in structural
geology and crystallography is covered elsewhere, particularly by
Phillips (1971). Simple representation of a pole and a plane is
illustrated in Fig. 6.4.
The engineering geologist is likely to be confronted with a large
number of discontinuity orientations from anyone survey, and will be
concerned with statistical studies of the concentration or grouping of
these data, rather than with one or two individual discontinuities.
Since the stereographic or Wulff net in Fig. 6.4 is not area-true, it is
usual to use, for large numbers of data, an area-true net, usually the
Lambert equal-area projection or Schmidt net. Printed versions of
this net 200 mm in diameter with a tracing paper overlay are used in
DIP
/N
DIRECTION

W~ GRE AT CIRCLE

STRIKE
DIRECTION
FULL EQUATGRIAL
DIP Pl ANE

(0 J ( b)
z'
N N
OLE PLANE

z
W .

CIN\ w~, "


@ • P'

S 21L· /50· DIP S

LOWER
\Y!fJ!JJ \:!lJ z' UPPER
LOWER HEMISPHERE HEMISPHERE
HEMISPHERE UPPER STEREOGRAM STEREOGRAM MERI DIO NAL NET 10" INTERVAlS-
PROJECTION HEMISPHERE OF POLE & PL ANE OF POLE & PL ANE AREAS INCREASING TOWAROS
r( ) PROJECTIO, CMJTSIOE
(d) ( e1
0 0
Figure 6.4 (a) The block diagram illustrates a plane striking at about 120 with a dip direction of 50°, usually expressed as 210°/50
dip. (b) To obtain a spherical projection, inscribe a sphere about 0 as centre. Z is the zenithal point, Z' is the nadir. The intersection of
the sphere by the plane is a great circle. The intersection ofthe line normal to the plane passing through 0 with the sphere is the pole to the
plane. (c) To obtain a stereographic projection of the plane,join all the points on the lower or upper semi-great circles respectively to the
zenith or nadir. The intersection of these lines with the equatorial plane is the stereographic projection of the plane. The stereographic
projection of the pole is the intersection of the line PZ or PZ' with the equatorial plane at P'. (d) A stereogram is the projection of the
plane or pole on an equatorial projection. (e) A stereo-net is afamity of stereograms, usually at 100 intervals. These may be meridional
(Wulff net) or adapted to equal area (Schmidt or Lambert net) where data are to be contoured.
150 Engineering Behaviour of Rocks

discontinuity analysis. The net comprises meridional great circles and


polar circles, spaced at 2° intervals in equal-area projection.
It should be noted that whilst the equal-area net eliminates areal
distortions which will affect the accuracy of any quantitative
assessment of grouped data, use of the equal-area net for stability
problems involving individual discontinuities will introduce equiva-
lent angular distortions. Ideally, therefore, the stereonet rather than
the equal-area net should be used for representing discontinuities.
This is not always convenient and the error is, in any case, probably
not significant in a very approximate approach to analysis.
In plotting discontinuity orientations only one-half of the great
circle is used to define the intersection of the discontinuity with the
reference sphere (Fig. 6.4), and only one of the poles need be plotted.
There has been controversy as to whether in engineering this should
be the upper or lower hemisphere projection. Structural geologists,
looking down into the earth at individual discontinuities, tend to use
lower hemisphere projections. Crystallographers, concerned with an
upper optical axis, use upper hemisphere projections. Engineering
geologists have had a tendency to use upper hemisphere projections
for multiple pole plots on the grounds that the pole is projected in the
same direction as the discontinuity dip, and lower hemisphere
projections for individual discontinuity plots which project the plane
of the discontinuity in the direction of the discontinuity dip. The
tendency now is to use lower hemisphere projections for both
purposes. It should be remembered that both projections can be
transformed by rotation through 180°. It is important that any
projections should contain, in a prominent place, information about
the hemisphere onto which discontinuities or poles are being
projected.
Preferred orientations of discontinuities from a joint survey can be
determined from a plot of poles to discontinuities on an equal-area
projection. An example of this is given in Fig. 6.5(a). It is important
when plotting these data to indicate major discontinuities. It is less
important to contour the zones of equal density distribution. The
method of contouring is outlined in Phillips (1971) and Hoek and
Bray (1974), and it is useful where all the discontinuities have equal
importance. In engineering, where major faults and shear zones have
exceptional importance, contouring can quite often hide or reduce the
importance of major discontinuities. The example in Fig. 6.5(b)
which contours plots of the discontinuities in Fig. 6.5(a) illustrates
this.
Discontinuities in Rock Masses 151

6.3 Shear resistance of a rock containing a discontinuity

The effect of a single discontinuity on the strength of a rock specimen


in uniaxial or triaxlal compression will be to reduce the strength of the
rock if the shear resistance of the discontinuity, usually expressed in
terms of the discontinuity friction angle, ¢d' is less than that of the
intact rock and it is inclined at an angle between ¢d and 90° to the
major principal stress. In Fig. 6.6, sliding will occur if the inclination
of the plane lies between f31 and f32'
The situation can be analysed if the shear resistance of the
discontinuity is represented by:
(6.1)
and the normal and shear stresses on a plane in the specimen whose
normal is inclined at an angle f) to the major principal stress axis are
represented by the Mohr circle defined in equations (2.28) and (2.29).
In order to retain an analogy with the intact material of Fig. 4.1, a
hypothetical Coulomb shear envelope for the rock is also included.
Then if the discontinuity is inclined at an angle f3 to the major
principal stress, the shear and normal stresses on the discontinuity
will be given by:
(6.2)

(6.3)
Note that effective stress notation is used throughout since it is
assumed that all fissures in rock masses may be filled with ground
water.
To obtain a modified strength criterion for the specimen containing
a discontinuity, equations (6.2) and (6.3) can be substituted in
equation (6.1) giving:
((1~ - (1~)(sin 2f3 - tan ¢d cos 2f3) = ((1~ + (1~)tan ¢d + 2Cd (6.4)
or
(1~[sin(2f3-¢)-sin ¢J-(1~[sin(2f3-¢)+sin ¢J= 2Cd cos ¢d
(6.5)
or

(6.6)
(al

, ,"
", ',

,," '
" ,
,'
, 0

,
o POLES TO FAUL S
OISCONTI UITIE5IUNSHEAREOI
(SHEARED I I LOWER HEMISPHERE PROJ EeTlON I
N
( b1

I·::::·";:} OVER 6%
( LOWER HEMISPHERE PROJ EC ION 1
8 - 16%

~ - 8%

D 2-~%
CONTOURS AT 2 L S 16% PER 1% AREA
77 READINGS

Figure 6.5 Plot of poles to discontinuities obtained from quarry scan-lines in


an exposure of Cambrian quartzites. Note that although the contoured plot (b )
indicates peak concentrations of discontinuities with azimuths and dips of
70°/70°,210°/35° and 310°/70°, these comprise unsheared bedding planes or
joints with a high level ofcontinuity. The most likely failure planes are indicated
in the uncontoured plot (a ) and are a fault at 310°/85° and three sheared
discontinuities at 130°/60°, all of which were open with a high degree of
continuity. (Data surveyed by R. D. Gavshon.)
154 Engineering Behaviour oj Rocks

1 1 :( · Ont tan ~ _ _~

'( d =Cd' Gnd tan <bd

0,

(T, On

Figure 6.6 Conditions for sliding along a discontinuity represented by a plane


whose normal is inclined at an angle p to the major principal stress (compare
with Fig. 4.1).

and
2 + 2(a;ICd )tan ljJd a;
---~--'---c----+- (6.7)
Cd (I-cot p tan ljJd)sin 2P Cd
The relation between a~ and P for tan ljJd =0.67 is illustrated in
Fig. 6.7 (a). This shows that sliding can occur in the range ljJ d < P< nl2
depending on the values of a; and a;, and that a;ICd will have a
minimum value when tan 2P= -cot ljJd' in this case for P=62°.
There will of course be a maximum value for a~ and this will be
given by equation (4.5) for the fracture of intact rock. If this is
computed for the case of Fig. 6.7(a) by putting ljJd = ljJ and C = 2Cd,
then the curves take the form of Fig. 6.7(b).
The shape of these curves is quite familiar and can be compared
with curves relating strength to anisotropy direction. Fig. 6.8
illustrates Attewell and Sandford's (1974) data on Penrhyn Slate and
Donath's (1964) test data on Martinsburg Slate. Typically the
specimen strength is at a maximum when a 1 is parallel or normal to
the cleavage planes. It has a minimum value when Pis about 30°. The
Discontinuities in Rock Masses 155

101 1 bl

10

Zd I
I
1_. . 3l 6'
:----__---:'30:' L __ 60 -1-1_ _ _-1-I_-----:LILI_ _ _:-:--_ _
90 3, (3 ~o
~I
90

Figure 6.7. Conditionsfor sliding along a discontinuity represented by a plane


whose normal is inclined at an angle p to the major principal stress for various
magnitudes of (11 and (13' normalized in terms of the discontinuity cohesion for
tan 4>d = 0.67. In plot (b) the curves are terminated by fracture of intact rock at
4> = 4>d' C = 2Cd·

implication of this type of curve is that sliding along the cleavage


planes significantly reduces specimen strength when they are
favourably orientated.
Stoney and Dhir (1977) have shown that three-dimensional
anisotropy can also be important. Fig. 6.9 illustrates this in the case of
Blair Atholl Marble where a three-dimensional plot of anisotropy
ratios for strength, 50% secant deformation modulus and axial strain
at maximum stress shows planar symmetry and variation between
planes. Based on the fabric, the x-axis is parallel to layering and
lineation, the y-axis is parallel to layering and normal to lineation and
the z-axis is normal to both layering and lineation. It is interesting to
note that in the xz and yz planes controlled by the layering the form is
similar to the typical transverse anisotropy curves of Fig. 6.8. In the
xy plane controlled by lineation, there is little variation in anisotropy
ratio. Of equal interest is the apparently greater sensitivity of
deformation modulus to anisotropy than strength.
The relation between the inclinations of a single discontinuity in a
INCLINAnON Of ll'iE CLEAVA(jE PLANE TOTHE
OIRECTION OF APPLICATION OF THE MA JOR
PRIN(IPAL STRESS 0 1

300 03=35 MN/mz

MARTINSBURG SLATE
'"!: 03 =10·5 M Iml
z 200
:>:
0"3=3.5 MN/mz
b'
b
100

0° 45° 60° 75' 90'

IN(LlNATION OF ANISOTROPY TO SPE(IME AXIS

Figure 6.8 Relation between deviatoric stress at failure and inclination of


cleavage (Penrhyn Slate) and anisotropy direction (Martinsburg Slate) to the
major principal stress or specimen axis direction (respectively, after Attewell
and Sandford 1974 and Donath 1964).
Discontinuities in Rock Masses 157

.- "
, MODULUS/STRE GTH RATIO
"..

COMPRESSIVE STRENGTH

Figure 6.9 Three-dimensional plot of modulus/strength ratio, compressive


strength, deformation modulus and strainfor Blair Atholl Marble, normalized to
the minimum value in each case (after Stoney and Dhir 1977).

two- or three-dimensional stress field and the magnitude of the


stresses can be studied by using the Mohr circle construction. The
values of [31 and [32 in Fig. 6.6 can be shown to be:

[31 =~+ ¢d _ ~ Sin-l((O"~ +0"3+ 2Cd cot ¢)sin ¢) (6.8)


2 2 2 O"~ - 0"3

[3 _¢d ~. _1((0"~+0"3+2Cdcot¢)Sin¢)
(6.9)
2-2+2 sln "
0"1-0"3

which can be rewritten:

(6.10)

(6.11 )
158 Engineering Behaviour of Rocks

In these equations it should be noted that the term Cd cot ¢d is


effectively the distance between the origin and the abscissa intercept
of the discontinuity shear resistance envelope 't"d = Cd +and tan o/d' If
this is added to a~ and a; it effectively moves the origin to the abscissa
intercept. In equations (6.8) to (6.11) this can be seen to be the case. It
is a neat way of allowing for discontinuity cohesion in a similar way to
water pressure in the concept of effective stress. It also means that
cohesion can be eliminated to simplify the mathematics of a three-
dimensional approach, without altering the basic principles.
So far only the major and minor principal stresses have been
considered. Jaeger and Cook (1969) show that the effect of the
intermediate principal stress on a discontinuity can be demonstrated
by considering the two extreme cases. These are the case implicitly
assumed previously where the minor principal stresses are equal
(a~ > a; = a;) and the other extreme case where the major principal
stresses are equal (a~ = a; > a;). In both cases the stresses may be
represented by a single Mohr circle. Then if values of /31 and /32 are
calculated from equations (6.10) and (6.11) for Cd=O, Ild=0.67, the
magnitudes given in Fig. 6.10 can be obtained for the ratio a3/a~.
In the case where O'~ > O'~ = O'~ sliding can occur on any plane whose
normal is at an angle between /31 and /32 to the 0'1 axis. The solid angle
represented by this is illustrated on the hemisphere and stereo graphic
projection in Fig. 6.1O(a). In the case where a~ =0';>0'; sliding can
occur on any plane whose normal is at an angle between tn - /31 and
tn - /32 with the 0'3 axis. The solid angle is represented by the
hemisphere and stereo graphic projection in Fig. 6.10(b). Jaeger and
Cook show that a dO' 2 values intermediate between these extreme
cases can be represented on a series of stereo graphic projections
grad ually changing the sliding windows from case (a) to case (b).

6.4 Shear resistance of a discontinuity

The terms Cd and ¢d have been used to define discontinuity shear


resistance in a very simple way in the previous analysis. The actual
shear resistance of a discontinuity is more complex and will depend
on various factors ranging from infill to surface properties and
continuity.
A useful starting point is to consider resistance to shear along a
single discontinuity, having frictional resistance but no cohesion. If a
shear box text (Fig. 6.11) is carried out along a single discontinuity at
0,
67 13, 13 ,
00 897 340
002 89 0 345
0<)5 880 360
010 855 JB 0
o5 825 4 0
o 10 785 450
025 130 51 0
o 30

0,

Figure 6.10 Effect of intermediate principal stress on discontinuity shear


stability illustrated through two extreme cases: (a) a; > a 2= a; and (b)
aj = a2 > a; where a; /a; = 0.20. The solid angle in which sliding can occur in
each case is shaded on the hemisphere and stereographic projection. Since
a;/a;=O.2, in (a) a;/a 2=0.2 and in (b) a;/a 2 = 1. Different magnitudes ofa 2
will see a change in solid angle from that in (a) to that in (b) (after Jaeger and
Cook 1969).
160 Engineering Behaviour of Rocks

E CAPSULATI G
MATERIAL
ROO< SPECIMEN

EST HORIZONTAL

SPECIME CARRIER

Figure 6.11 Suggested arrangement for a laboratory direct shear test on a


single discontinuity (after Brown 1981).

a low normal stress, then the shear stress-displacement curves for


tests on smooth and rough discontinuities in the same rock, and at the
same constant normal stress, will be similar to those outlined in
Fig. 6.12. During the test on the rough surface (a), shear resistance
will rise with displacement to a peak shear stress. During the test there
will be dilation as the asperities which make up the surface roughness
are mounted. As movement takes place on the downside of the
discontinuities, the shear resistance will then fall to a residual level,
rising again as the next series of asperities are mounted.
If the normal stress is increased - and to a lesser extent if the
roughness of the discontinuity is reduced - this pattern of behaviour
will change. Initially the asperities will still be mounted, but the larger
discontinuities will tend to be sheared at high strains, giving the type
of modified strain-softening behaviour in Fig. 6.12(b). This tendency
will increase with increasing normal stress until a typical strain-
softening curve as in Fig. 6.12(c) is obtained. Ultimately at very high
normal stresses fracturing through the discontinuity will occur
immediately and there will be no mounting of discontinuities, leading
to a strain-hardening type of behaviour as in Fig. 6.l2(d).
Ifthe hypothetical curves from Fig. 6.12 are compared with actual
stress-strain curves in Figs 4.13 to 4.19 for intact rock under triaxial
test conditions, similarities can immediately be seen. Both tests
demonstrate a change from poorly controlled, strongly dilatant
strain-softening behaviour through a brittle-ductile transition to
stable, mildly dilatant, strain-hardening behaviour. This similarity
has been used by Einstein and coworkers (see Einstein and
Discontinuities in Rock Masses 161

_ -- -- -- - - Idl HIGH
ORMAL
STRESS

- - - - - - - - - - - leIIN(REASI G
NORMAL
STRESS

( b I MODER ATE
NORMA L
STRESS

I Q I LOW NORMAL STRESS

lei
f SHEAR
10)

( ORMAL

Figure 6.12 Discontinuity shear resistance against shear strain, and normal
strain against shear strain during shear testing along a typical discontinuity
(after Roberds and Einstein 1978).

Hirschfield 1973; Roberds and Einstein 1978) to develop a


comprehensive model for rock discontinuities on the lines of the
critical state model outlined in Chapter 4. Before considering this
approach it is, however, useful to consider some of the other models
for discontinuity behaviour.
Consider for instance a series of tests carried out at low normal
stresses on two types of discontinuity. The first is smooth with no
asperities. The second has asperities, but the normal stresses an:
sufficiently low to avoid shearing of asperities.
Then if a series of tests at increasing normal stresses are carried out
and the results plotted as shear resistance at normal stress for both the
peak shear resistance of the rough and smooth surfaces, there will be
two approximate linear relationships, with the equations:
(6.12)
162 Engineering Behaviour of Rocks

for the rough discontinuities, where Cd is the additional initial


resistance required to mount the discontinuity, and
(6.13)
where <Pr may be considered the residual or absolute friction angle of
the surface without discontinuities, although it would be difficult to
imagine a perfectly smooth surface.
The effect of asperities has been considered by Patton (1966) in a
general model of discontinuity shear resistance. The basic geometry is
illustrated in Fig. 6.13. This models the asperities as planes inclined at

1--- - >-

Figure 6.13 Model of a discontinuity surface as a series of regular asperities


(after Patton 1966).

an angle i to the direction or plane of sliding. Then the equation for


sliding on the asperity planes will be
(6.14 )
and the shear and normal stresses acting in the direction of sliding, !d
and O"~d' giving the equation for sliding along a rough discontinuity,
will be related to ! i and O"~, by:

!d !i cos i+O"~i sin i A,.


-,-=, . . . =tan 'I'd (6.15)
O"nd O"niCOS l-!,Slll I
whence
(6.16)
and for the case of peak shear resistance due to mounting an asperity:
(6.17)
Discontinuities in Rock Masses 163

Thus for a series of regular projections (Fig. 6.13) the dilatant part
of the shear resistance envelope for a rough plane discontinuity can be
expressed in terms of equation (6.17), where i = tan -1(2A/Je) with A as
the asperity dimension and Je the wavelength.
This equation was confirmed in model tests by Patton, but more
importantly Patton and Deere (1970) demonstrated its practical
application. In particular, they demonstrated from a series of
observations of bedding planes controlling the stability of unstable
limestone slopes that the critical inclination of the bedding plane was
given by the sum of the average of i for the plane and ¢r determined
from laboratory tests. For the same rock, stable bedding plane
inclinations varied between 31 0 (smooth planes) and 60° (rough
planes).
Some typical values for ¢r are given in Table 6.1 together with some
typical discontinuity infill ¢ values. The lower values are generally for
wet surfaces. Goodman (1976) suggests that, for all except micaceous
rocks, tan ¢r should lie between 0.5 and 0.6.
In fact Patton's results were rather conservative, since the low
normal stresses involved made it necessary only to consider first-
order i angles. Where interlocking or high normal stresses prevent
dilation, fracturing of surface projections is an important factor
determining the strength of discontinuities. In order to allow for this,
Patton's model has been modified by Ladanyi and Archambault
(1969) and by Barton (1973).
Ladanyi and Archambault combined the contributions offriction,
dilation and interlocking to shear strength in the form:
I (1-as)(d£vld£s Han ¢r)+as'f
'd=(Jnd l-(l-as)(d£v/d£s)tan¢r (6.18)

Table 6.1 Typical smooth discontinuity <Pr values and infill material <P
values (Hoek 1970)

<Pr (deg) Infill <P (deg)

Granite 31-33 Clay gouge 10-20


Limestone 33-37 Calcite 20-27
Sandstone 25-34 Breccia 22-30
Quartzite 26-34 Rock aggregate 40
Shale 27-32
164 Engineering Behaviour of Rocks

where as is the proportion of total discontinuity area sheared through


asperities, dey/des is the dilation rate at peak shear strength and is 'f
the strength of the intact rock material. Equation (6.18) reduces to
equation (6.17) when as =0 and dey/des =tan i and to equation (6.14)
when as = 1 and, = 'f' The equation is therefore a curve asymptotic to
the two curves represented by these equations in Fig. 6.14 and allows
for mounting of asperities at low normal stresses and shearing of
asperities at higher normal stresses.

ROUGH SURFACE ~ ROUGH SURFACE


UNSHEARED -SHE ARED ASPER ITI ES
ASPERITIES ld : Cd + an"d to n <b r
'Cd = and ton l<b r + d

Figure 6.14 Representation of shearing envelopes for discontinuities at


different normal stresses.

A more readily applicable approach is Barton's empirical shear


strength equation

'd = (J~d tan[J RC IOg10(~~:)+ 4>rJ (6.19)

where J RC represents a joint roughness coefficient having a sliding


scale of roughness varying from 0 in perfectly smooth rock to 20 in
very rough rock, JSC is the effective joint wall compressive strength
and 4>r is the residual friction angle.
An illustration ofthe method is given in Fig. 6.15 for JRC values of
20, 10 and 5, J CS values of 5, 10, 50 and 100 MN m - 2 and 4>r = 30°.
Discontinuities in Rock Masses 165

ROUGH UNDULATING - SMOOTH UNDULATI G- SOOTH NEARLY PLA AR-


TENSIO N JOINTS .I1OUGH SMOOTH SHEETING.NON·PlANAR Pl ANAR SHE.AR JOINTS
SHEETING. ROUGH BEDDI NG FOLIATION' "rooING. PLANAR FOLIATION. BEOOING
50 em SOem 50 em

E 500cm
-1 500 em SOOcm

E E
'T1(fn= ton[ 20 loglO (~) • 30·] Tlern' ton [10 . I0910 (!;;) .30°] TI(fn= tan [5 10910 (~) • 30°]
I A) f BI fel

..
~ f--t----,>f--"'71---;
5

I.
I
1:

131---u<---;""""'--;
'"....~ I-H--¥~<::":"-; /-
~J
::.:~
:5
~ I+h"'-+--t--;:--;
~~
'/
po' (
JRC,S
o o 1

NORMAL STRESS MN/ m2 ORMAL STRESS N/m2 NORMAL STRESS M 1m2

Figure 6.15 Illustration of Barton et a/"s (1974) empirical relation between


discontinuity shear resistance and normal stress for discontinuities of varying
roughness. The numbers on the curves represent JCS in MN m- 2 .

The range of tPd values (tan-l(!d/(j~d)) between 30° and 60° may be
low at the top end and Barton suggests a linear portion equivalent to
!d/(j~d = tan - 1 (70°) for low normal stresses in strong rocks with
rough discontinuity surfaces.

6.5 A critical state model for rock discontinuity strength

In addition to the three models discussed, other models for


discontinuity shear have been suggested by Goodman (1974),
Ghaboussi et al. (1973) and Roberds and Einstein (1978). The latter is
particularly interesting in its similarity to the critical state models for
rock deformation discussed briefly in Chapters 4 and 5.
Roberds and Einstein (1978) take the four types of discontinuity
stress-strain curves in Fig. 6.12 and characterize the shear stress-
strain paths in terms of the stress ratio, rd/(j~d = tan tP~. Then the
change in shear mechanism from mounting of asperities to shearing
166 Engineering Behaviour of Rocks

tan cP YIELD STRAIN


CRITICAL
SOFTENING
STATE
Ian 1>r

POSSIBLE
CO MPACTIO &
COMPRESSION

' ·0
..-
an
SHEARING THROUGH INTACT
MOUNTI NG I ASPERI IES ROCK
OF
ASPER ITIES I
Hn I DlLATIO
~

Figure 6.16 A general model for discontinuity shear resistance related to the
critical state (after Roberds and Einstein 1978).

through of asperities can be expressed in terms of the stress ratio as a


change from tan cP;ield to tan cPr' or as a ratio tan cP;ield/tan cPr. In Fig.
6.12 tan cPyield will represent the peak stress ratio in (a), (b), (c) and
tan cPr the average residual stress ratio in (a) and (b) and the residual
stress ratio in (c). Tan cPr is the stress ratio for a completely smooth
surface, with all the asperities sheared off, so when in (c) or (d)
tan cPyield = tan cPr' unlimited shear will occur with no dilation. This
represents the ultimate residual stress ratio or the critical state.
Both the reduction in the ratio between stress ratio at yield and
residual stress ratio (tan cPyield/tan cPr) and the dilation rate from the
flow rule (dsnormal!dsshear), are stress-dependent and a model for the
yield condition and flow rule for a discontinuity can be illustrated in
Fig. 6.16.
Discontinuities in Rock Masses 167

6.6 Measurement of discontinuity shear resistance

Various methods have been suggested for measurement of discon-


tinuity properties. These range from large-scale in situ tests on exposed
blocks to laboratory shear box tests (Fig. 6.11) on discontinuities
exposed in core samples. Various methods are described in Hoek and
Bray (1974), Goodman (1976) and Attewell and Farmer (1976). Some
tests are useful, some are less useful, and the present book is not the
correct place for a critical appraisaL It is sufficient to quote Goodman
that: 'no laboratory or field test, however carefully controlled can
duplicate the scale and character of the loading, boundary and
environmental conditions inherent to engineering service'.
The only laboratory test which can be performed satisfactorily and
easily is that to obtain ¢;, which can be obtained through residual
shear tests on flat sand-blasted or even rough sawn rock surfaces in a
shear box. The test should ideally be carried out with the plane
completely immersed in water, although for comparison it may also
be carried out with the surfaces dry. Typical ¢; values are quoted in
Table 6.1. Infill material properties can be estimated using conven-
tional soil mechanics tests.
The best way in which actual field ¢d values can be obtained
satisfactorily is on the basis of the estimating procedures quoted in
previous sections, or by back-analysis of existing similar case
histories. Hoek and Bray (1974) use the case history approach to
estimate Cd values for various rock types. In stability analysis it is
often easier to assume Cd =0 and estimate ¢d. Cohesion is as
unsatisfactory a concept in fissured rock as it is in soils. It depends on
too many factors including discontinuity roughness, normal stress,
moisture content, width and infill of fissures, type oftest and degree of
weathering to inspire confidence as a quoted material characteristic.
7 Behaviour of
Rock Masses

Deformation of sufficient magnitude to constitute failure of massive


rock, particularly near to the surface, usually results from movement
along a single discontinuity or a combination of discontinuities.
Description of the orientation, dip and surface characteristics of
major discontinuities and discontinuity sets in a rock mass normally
provides sufficient information to design a structure in a massive
rock. There remains, however, a demand for a more general
description of the behaviour of rock masses which has resulted in
considerable research into mainly empirical methods for describing
rock mass behaviour.
Most of these derive from Terzaghi's (1946) classification of rock
masses which has been described in Chapter 1 (Tables 1.6 and 1.7)
and which was the first attempt to relate a systematic description of
rock to its mechanical reactions - in this case for tunnel arch support
loads. Other similar classifications which have been developed for
tunnels, foundations and slopes have also been mentioned in
Chapter 1, as have Deere's (1964) concept of rock quality designation
(RQD) and Onodera's (1963) concept of rock mass factor U). The
importance of these latter concepts is that they provide a method of
relating the behaviour of the rock mass - containing discontinuities,
water and zones of weathering - to the behaviour of an intact rock
specimen.
This is an ideal for which engineers designing in rock are
increasingly searching. In this final chapter, some of the approaches
used and their implications are considered.

168
Behaviour of Rock Masses 169

7.1 Discontinuity frequency

In any attempt to assess the behaviour of massive rocks, the frequency


of occurrence of discontinuities and the strength of the intact rock
between them becomes as important as the characteristics of an
individual discontinuity. In borehole logging, discontinuity fre-
quency is sometimes characterized as a fracture index, defined as the
number of fractures per metre length of solid core recovered. Deere
(1968) has also proposed a classification based on discontinuity
spacing, summarized in Table 7.1. It is worth comparing this with
Deere's RQD classification in Table 1.9 and Terzaghi's classification
in Table 1.6, and noting that RQD is used as a description for the more
discontinuous rock and also that neither classification allows for
Terzaghi's hard stratified or schistose rock. This may have numerous
laminar partings if drilled normal to the bedding, but still remain a
high-quality 'intact' rock for most engineering purposes.

Table 7.1 Discontinuity spacing classification (after Deere 1968)

Discontinuity Fracture Terzaghi


spacing index classi-
Rock mass (m) (m- 1 ) fication

Solid >3 1
Massive 1~3 <1 3
Blocky/seamy OJ~1 1~3 4~5
Fractured 0.05-DJ 3~20 6
Crushed/shattered <0.05 >20 7

Whilst discussing discontinuity frequency, it is worth noting the


observation of Rocha (1976), Snow (1968) and Londe (1973), among
others, that discontinuity spacing distributions in rocks tend to follow
a negative exponential distribution. Priest and Hudson (1976) have
shown that this can take the form (Fig. 7.1):
(7.1 )
where x is a discontinuity spacing value,J(x) is the probability density
distribution of x and .Ie is the mean number of discontinuities per
metre.
In this case, the probability of discontinuity spacings occurring
TOTAL SCANLINE LE GTH 51457m
~~ 11 MEAN DISCONTI NUITY SPACI G (i I o 105m LOwE R
::> -'
Zu
10 STANDARD DEVI ATION o 13m CH ALK
~ J: 9 UMBER OF ALUES 48e4
~:.t 8
~UJ 7
15~ 6
~'" S
.... ~ 4
'" :;l: FITTED NffiATiVE EXPONENTIA
~ '" 3 PROBABILITY DE SHY DISTRIBUTION
3<!:
w u
2 1= 9-488 / m >0 -50
~a 0lUJlLUUlllUlllWU~~~PU~~~~~~~~~~
~ V1
001 0 2S 0 30 0 35 0-40 O· 5 0 0
UI TY SPACING m

TOTAL SCAN LINE LENGTH 32 22m CARBO IFEROUS


MEAN OISCONTI UITY SPACI GIKI 0 ·129 SA DSTONE
STANDARD DEVIATION O· 16m
UMBER OF VAWES 249

FlmD NffiATIV't EXPONE TIAL


PROBABI LI TY DENSITY OISTRIBUTION
1 : 1. 728/m

>050 ,

040 0-45 0 SO

TOTAL SCANU E LENGTH 60 -73m


MEA DISCO TI UITY SPAC ING (il 0033
STANOARD DEVI ATION 0-032m CARBON IFERO US
NUMBER OF VALLES 1828 MUDSTO NE

FITTED NEGATIVE EXPONENTIAL


PROBABILITY DE SITY DISTRI BUTiO N 1: 30 .121m

0·01 0-03 ~OS 0·07 0 ,0'1 0," 0 ·13 0 -\5


OISCONTlNUI Y SPACING ,m_

Figure 7.1 Discontinuity spacing histograms from tunnel scan-lines logged by


Priest and Hudson (1976), illustrating the negative exponential distribution of
discontinuit y spacings in rock.
Behaviour of Rock Masses 171
between x and x + dx is given by f(x) dx; and if the total number of
discontinuities along a scan-line of length L is given by )"L, then the
total number of intact lengths between x and x + dx is )..Lf(x) dx and
the length of these is )..Lxf(x) dx. Then the length ofthe ith core will
be:
x;=)"Lxf(x) dx (7.2)
Equation (1.20) defines RQD in terms of the ith core length and if
equation (7.1) is substituted in equation (7.2) and then substituted in
equation (1.20), an expression for RQD in terms of ).. and x can be
obtained as dx -+0 in the form:

RQD= 100)..2 (L xe- AX dx (7.3)


JO.1
for the arbitrary threshold value of 0.1 m.
Then, if L is large and e-AL-+O this solves to give:
RQD= 100(0.IA+ l)e- o.1A (7.4)
Comparison of measured and theoretical RQD values, in Fig. 7.2,
the latter computed by Priest and Hudson from Deere's data, shows
the applicability of this approach, although at low RQD the
theoretical value tends to be unrealistically high. The reason for this is
that at low values the condition of the rock and the width of

...
100 ' -- ""'Tl£OI!==E="-=-CA""L--' ~Or-~~TH~EO~R~ET~D~l'-'

i' ,
75 .": CU~VE RJOO . CURVE ROO -
75
~
§ 50 • .:..
,. 'if!. so
Cl
a:
25 •
a,N X CORE HAO<ENSTOCK
g 25
o SIL~TONE
IblNX CORE
JOHN DAY BA SALT
o2 4 6 8 10 1214161820 00 2 4 6 8 10 12 14 16 18 ZO

100 ro.-r.:---=:-;;== ,--, 100.--fl-rr.,...-::====c---.

75 75
o TUNNEL WA LL SCAN lINE
~ 50 ~ so ACROSS OISCONTl NUITIES .
Cl
o
g 6 TU NNEL WALL SCANLINE
£r Z5 000 PARAL LEL TO OISCONTlNUInES
cr 25 , NX CORE OWORSHAK DAM
o NX CORE .
G~ANlTE GNEISS 10,ClIMAX STOCK G~NI~E
o0 2 4 6 8 10 12 1416 18 20 o0 2 4 6e 10 12 14 16 1
A.VERAGE NUMBER OF DISCONTINUITIES' PER m ~

Figure 7.2 Comparison of measured and calculated RQD values by Priest and
Hudson (1976) from discontinuity frequency data collected by Deere (1968).
172 Engineering Behaviour of Rocks

discontinuity openings - indicators of weathering - will have a


significant effect on measured RQD.

7.2 Rock mass classification systems

The need to provide an index of massive rock behaviour which would


include allowances for discontinuity frequency, discontinuity orien-
tation, discontinuity surface properties, intact rock strength and
geostatic and hydrostatic stress effects led, during the 1970s, to the
development of several quite complex rock mass classification
systems. The best-known are the Rock Structure Rating (RSR) system
of the US Bureau of Mines (Wickham et al. 1972), the Geomechanics
Classification of the Council for Scientific and Industrial Research,
South Africa (Bieniawski 1973) and the Rock Mass Quality (Q) system
of the Norwegian Geotechnical Institute (Barton et al. 1974).
The purpose of these classification systems is to isolate in a
qualitative manner categories of rock which, under a particular set of
engineering constraints, will behave in a similar way. The principle on
which each is based is that a number of parameters are selected which
can be determined quickly and easily by a field geologist or engineer
and which are important in determining rock mass behaviour. For a
particular rock, a numerical assessment is made within designated
limits of the quality of the rock as it relates to a particular parameter.
These numbers are then summed or multiplied to give an index
number which can be related to the behaviour of the rock mass.
The RSR system was developed to predict support densities for
tunnel drives, from a study of the performance of33 tunnels in various
strata. Three parameters, A, Band C, defining the general geology, the
joint pattern related to the drive direction and the groundwater and
joint condition, are selected and allocated the values in Table 7.2. The
rating is then obtained as:

RSR = (A + B)+ C (7.5)

The parameters are self-explanatory, and it may be noted that most


are very simple and well defined. The most remarkable omission is
rock strength, the basic assumption being that, at the depths
encountered, deformation is unlikely to result from intact rock
fracture. It is an assumption which is inherent in many analyses of
rock mass behaviour from Terzaghi (1946) to Hoek and Brown
(1980b). The use of the system is illustrated in Fig. 7.3, where RSR
Table 7.2 Rock Structure Rating system parameters (after Wickham et al.
1972)
(a) Rock structure rating parameter A. General area geology. Max. value 30

Basic rock type Geological structure

De-
Hard Med. Soft compo
Slightly Moder- In-
Igneous 1 2 3 4 faulted ately tensely
Metamorphic 1 2 3 4 Massive or faulted faulted
Sedimentary 2 3 4 4 folded or or
folded folded

Type 1 30 22 15 9
Type 2 27 20 13 8
Type 3 24 18 12 7
Type 4 19 15 10 6

(b) Rock structure rating parameter B. Joint pattern. Direction of drive. Max.
value 45

Strike .1 to axis Strike II to axis

~
Direction of drive Direction of drive
-56
~ 48
-540
~32 Both With dip Against dip Both
g'24
U 16 4

" Dip of prominent Dip of prominent


~ 8 23
0 joints· joints·
o 8 162432404856
Thickness (inches)
Dip- Verti- Dip- Verti- Dip- Verti-
Flat ping cal ping cal Flat ping cal

1 Very closely jointed 9 11 13 10 12 9 9 7


2 Closely jointed 13 16 19 15 17 14 14 11
3 Moderately jointed 23 24 28 19 22 23 23 19
4 Moderate to blocky 30 32 36 25 28 30 28 24
5 Blocky to massive 36 38 40 33 35 36 34 28
6 Massive 40 43 45 37 40 40 38 34

* Notes: flat =0-20°; dipping = 20-50°; vertical = 50-90°.


174 Engineering Behaviour of Rocks
Table 7.2-continued

(c) Rock structure rating parameter C. Ground water. Joint condition. Max.
value 25

Sum of parameters A + B

Anticipated 13--44 45-75


water
inflow
Joint conditiont Joint conditiont
(GPM/lOOO ft)

Good Fair Poor Good Fair Poor

None 22 18 12 25 22 18
Slight ( < 200 GPM) 19 15 9 23 19 14
Moderate
(200-1000 GPM) 15 11 7 21 16 12
Heavy (> 1000 GPM) 10 8 6 18 14 10

t Notes: good = tight or cemented; fair = slightly weathered or altered; poor = severely
weathered, altered or open.

numbers are related to rock loads and support systems for a 4.2 m
diameter tunnel formed by drill and blast. Because of the complexity
of the empirical relations, imperial units have been retained in this
figure and in Table 7.2.
The Geomechanics Classification in its modified form (Bieniawski
1976) is based on five parameters, rock strength, RQD, discontinuity
spacing and condition and groundwater conditions. The numerical
assessment rating of each of the parameters is given in Table 7.3(a)
together with adjustments for joint orientations in Table 7.3(b). The
classification rating (sometimes called rock mass rating, RMR) is
obtained by summing the individual parameter ratings:
RMR=(1 +2+3+4+5)-(b) (7.6)
In Table 7J(c) the suggested classification based on the rating is
summarized and comments on the meaning of the classes are included
in Table 7J(d). It is interesting to note in Class I the reduction in
intact strength from 200 MN m- 2 to OJ MN m- 2 , effectively
suggesting a reduction in strength from peak to residual when
considering rock masses. This has been discussed in Chapter 4. It is
not wholly compatible with the estimated stand-up time for a 5 m
Behaviour of Rock Masses 175

::
er
0 '"'
<t"U
oC:
.J::J
<>
c:.
,-,0

70
0:=
o-=<

0)
I ~ ~_ ~ -
_- RIB

--
60
10 :b ~ !,SJ- SECTIO
~
z
~ 50 I
a: I b ~ 1>_
I .. " ..... '
I

I PRACTICAL
1-0IAMET~R I LIMIT FO~
lO ROCK BOl TS I 1118 AND
m- I__ BOLT
S 'J;;;; I SPACING
Wr I
~!..v'! ___ L ___ _
I rlu rD 10, '00' ,
I
10~~---L---L--~--L-~~~~~-

012345678
RIB SPACINGIf ) BOLT SPAC ,NG I It. f I SHDTtREli: THICI(I<ESS ( .nI

Figure 7.3 Illustration of the Rock Structure Rating system used to determine
rib spacing for various arch sections, bolt spacing and shotcrete thickness (after
Wickham et al.1972). Note that, in this figure and in Table 7.2, because of the
complexities of the empirical relations, units are retained in the original
Imperial, rather than SI, notation.

opening of 10 years. The concept of stand-up time - the average time


taken for an unsupported rock span to collapse - developed by
Lauffer (1958), is a particularly useful way of categorizing massive
rock and has been used by Bieniawski (1976) and Barton (1976b) to
translate classification ratings into meaningful engineering criteria.
Fig. 7.4, which includes Bieniawski's and Barton's classifications on
Lauffer's stand-up time chart, illustrates the really quite resilient
nature of unsupported spans, particularly over a limited time period.
The Rock Mass Quality system is more complex than the other
two, and the reasoning behind it is more cogent. It is based on analysis
of a large number of Scandinavian tunnel support case histories
collected by Cecil (1970) and is based on six parameters; RQD,
number of joint or discontinuity sets, joint roughness, joint alteration,
water flow and a stress reduction factor, based on the stress acting on
competent rock or the degree of weakness or disturbance in less-
competent or sheared rocks. The ratings for each of these parameters
Table 7.3 Geomechanics Classification of jointed rock masses (after Bieniawski 1976)
(a) Classification parameters and their ratings

Parameter Ranges of values

Strength Pomt load >8 4-8 2-4 1-2 For this low range,
of strength mdex umaxIaI compres-
mtact rock { SIve test IS preferred
materIal Umaxlal >200 100-200 50-100 25-50 10-25 3-10 1-3
(MN m-') compressive
strength
Ratmg 15 12 4 2 o
DrIll core quality RQD (%) 90-100 75-90 50-75 25-50 <25
Ratmg 20 17 13 8 3

Spacmg of Jomts >3m 1-3 m 0.3-1 m 50-300 mm <50mm


Ratmg 30 25 20 10

4 ConditIon of Jomts Very rough surfaces Slightly rough surfaces Slightly rough surfaces SlickensIded surfaces Soft gouge > 5 mm thick
Not continuous Separation < 1 mm Separation < 1 mm OR OR
No separatIOn Hard Jomt wall rock Soft Jomt wall rock Gouge < 5 mm thIck JOtnts open> 5 mm
Hard Jomt wall rock OR Contmuous Jomts
Jomts open 1-5 mm
Contmuous Jomts
Ratmg 25 20 12 6 o
Inflow per 10 m None <25 25-125 > 125
tunnel length
(I mm-') OR OR OR OR
Ground
Jomt water
water
pressure o 00-02 02-05 >05
Ratio major prIncipal
stress OR OR OR OR
General conditIons Completely dry MOIst only Water under moderate Severe
(mterstltIal water) pressure water problems
Ratmg 10 7 4 o
Table 7.3-continued
(b) Rating adjustment for joint orientations

Strike and dip Very Very


OrientatIOns of Jomts favourable Favourable Farr Unfavourable unfavourable

{TUnnels 0 -2 -5 -10 -12


Ratmgs FoundatIOns 0 -2 -7 -15 -25
Slopes 0 -5 -25 -50 -60

(c) Rock mass classes determined from total ratings

Ratmg 100-81 8G--61 60-41 40-21 <20


Class No II III IV V
Descnptlon Very good rock Good rock Fau rock Poor rock Very poor rock

(d) Meaning of rock mass classes

Class No. II III IV V


Average stand-up tIme 10 years for 5 m span 6 months for 4 m span 1 week for 3 m span 5 hours for 15m span 10 mmutes for 05 m
span
CohesIOn of the rock mass >300 200-300 150-200 100-150 < 100
(kN m- 2 )
FnctlOn angle of the rock mass (deg) >45 40-45 35-40 30-35 <30
178 Engineering Behaviour of Rocks

70 10min lhoor lOhoors 1da 1",eek 1manltl 6month1 ear 5 10 20 yea r s

~~~K~
Ey~~----~--+-----r---+---~ :r"--
" -I-~U 50
IIJ SO~H AFRI CAN 01+_-+----1---1-- .A""-I---+-+-; 40
30 f_+-A
- U-ST- R-'
AN, ...-- - -e
- - t- --!- __-L~~~~~~~~~~ 30

Q> 100
~1 5 f-+-----r,-
M M~E~
O'-A~
TE+--+--~~TIr~~
.. COL L APSE

~10 ~~~~~~~~~~
10

ct a ~ 8
~ 6 f_+-----r----,.q=,...n e 6
::: 5 f_+- - - I -- -7'---Y 5
:5
0-
4
~ 3 H------i>''-T,
VI
%
:> e

STAND UP TIME. hours

Figure 7.4 Comparison of the Geomechanics (South African) and Rock Mass
Quality (Scandinavian) methods of rock classification with Lauffer's (1958)
estimates of stand-up time and unsupported span based on observations in
Austrian alpine tunnels (after Bieniawski 1976).

are included in Table 7.4, and the rock mass quality is defined by:

Q=(RQD)(Jr)(~) (7.7)
In Ja SRF
The reasoning behind this equation is that each of the quotients
represents an important quantity of the rock mass. Thus RQDjJn
represents the structure of the mass and is a crude measure of block or
particle size. Barton et al. (1974) suggest that the range of values from
200 to 0.5 represents a slightly truncated size range for rocks if
expressed in centimetres. Jr/ Ja represents interblock shear resistance or
friction coefficient. The range of values is from 5 to about 0.1 and
represents the range from strongly dilatant rock to the residual
frictional coefficient of clay infill. The final quotient Jw/SRF can be
said to represent the active stresses, but it is complex, because SRF
represents different stress parameters, depending on the continuity of
the rock. Thus it is the loosening load when the excavation is through
Table 7.4 Ratings for Rock Mass Quality classification (after Barton et
al. 1974)

1. Rock quality designation (RQD)

A. Very poor 0-25


B. Poor 25-50
C. Fair 50-75
D. Good 75-90
E. Excellent 90-100

Notes' (i) Where RQD is reported or measured as ~ 10 (including 0), a nominal value
of 10 is used to evaluate Q.
(n) RQD intervals of 5, I.e. 100,95,90, etc., are suffiCiently accurate.

2. Joint set number (J n)

A. Massive, no or few joints 0.5-1.0


B. One joint set 2
C. One joint set plus random 3
D. Two joint sets 4
E. Two joint sets plus random 6
F. Three joint sets 9
G. Three joint sets plus random 12
H. Four or more joint sets, random, heavily jointed, 'sugar-
cube', etc. 15
1. Crushed rock, earthlike 20

Notes (I) For intersectIOns, use (3.0 x I n ).


(ii) For portals, use (2.0 x J n)'

3. Joint roughness number (Jr)


(a) Rock wall contact and
(b) Rock wall contact before 10 cm shear

A. Discontinuous joints 4
B. Rough or irregular, undulating 3
C. Smooth, undulating 2
D. Slickensided, undulating 1.5
E. Rough or irregular, planar 1.5
F. Smooth, planar 1.0
G. Slickensided, planar 0.5

Note' (I) DescnptlOns refer to small-scale features and mtermedlate-scale features, m


that order.
Table 7.4-continued
(c) No rock wall contact when sheared

H. Zone containing clay minerals thick enough to prevent


rock wall contact 1.0
1. Sandy, gravelly or crushed zone thick enough to prevent
rock wall contact 1.0

Notes: (ii) Add 1.0 if the mean spacmg of the relevant jomt set is greater than 3 m.
(iii) J,=0.5 can be used for planar slickensided joints having lineations,
provided the lineations are orientated for minimum strength.
4. Joint alteration number (Ja)
(a) Rock wall contact

Ja ¢, approx. (deg)

A. Tightly healed, hard, non-softening, im-


permeable filling, i.e. quartz or epidote 0.75
B. Unaltered joint walls, surface staining
only 1.0 25-35
C. Slightly altered joint walls. Non-softening
mineral coatings, sandy particles, clay-free
disintegrated rock, etc. 2.0 25-30
D. Silty-, or sandy-clay coatings, small clay
fraction (non-soft) 3.0 20-25
E. Softening or low friction clay mineral
coatings, i.e. kaolinite or mica. Also
chlorite, talc, gypsum, graphite, etc., and
small quantities of swelling clays 4.0 8-16

(b) Rock wall contact before 10 cm shear

F. Sandy particles, clay-free disintegrated


rock etc. 4.0 25-30
G. Strongly overconsolidated non-softening
clay mineral fillings (continuous, but
< 5 mm thickness) 6.0 16-24
H. Medium or low overconsolidation, soften-
ing, clay mineral fillings (continuous but
< 5 mm thickness) 8.0 12-16
1. Swelling-clay fillIngs, i.e. montmorillonite
(continuous, but < 5 mm thickness).
Value of Ja depends on percentage of
swelling clay-size particles, and access to
water, etc. 8-12 6-12
Table 7.4-continued
(c) No rock wall contact when sheared

K.{zones or bands of disintegrated 6, 8


L. or crushed rock and clay (see or
M. G, H, J for description of clay condition) 8-12 6-24
N. Zones or bands of silty- or sandy-clay,
small clay fraction (non-softening) 5.0
O.{ThiCk, continuous zones or
P. bands of clay (see G, H, J for 10, 13 or
R. description of clay condition) 13-20 6-24

5. Joint water reduction factor (Jw)

Water pressure
approx.
Jw (MN m- 2 )

A. Dry excavations or minor inflow, i.e.


< 5 I min - 1 locally 1.0 < 0.1
B. Medium inflow or pressure, occasional
outwash of joint fillings 0.66 0.1-0.25
C. Large inflow or high pressure, consider-
petent rock with unfilled joints 0.5 0.25-1.0
D. Large inflow or high pressure, consider-
able outwash of joint fillings 0.33 0.25-1.0
E. Exceptionally high inflow or water press-
ure at blasting, decaying with time 0.2-0.1 > 1.0
F. Exceptionally high inflow or water press-
ure continuing without noticeable decay 0.1-0.05 > 1.0

Notes: (I) Factors C to F are crude estimates. Increase Jw If dramage measures


are mstalled.
(ii) Special problems caused by Ice formation are not considered

6. Stress reduction factor (SRF)


(a) Weakness zones intersecting excavation, which may cause loosening
of rock mass when tunnel is excavated

A. Multiple occurrences of weakness zones containing clay


or chemically disintegrated rock, very loose surrounding
rock (any depth) 10
B. Single weakness zones containing clay or chemically
disintegrated rock (depth of excavation :!( 50 m) 5
Table 7.4-continued

C. Single weakness zones containing clay or chemically


disintegrated rock (depth of excavation> 50 m) 2.5
D. Multiple shear zones in competent rock (clay-free), loose
surrounding rock (any depth) 7.5
E. Single shear zones in competent rock (clay-free) (depth of
excavation ~ 50 m) 5.0
F. Single shear zones in competent rock (clay-free) (depth of
excavation> 50 m) 2.5
G. Loose open joints, heavily jointed or 'sugar-cube', etc.
(any depth) 5.0

Note: (i) Reduce these values of SRF by 25-50% If the relevant shear zones only
mftuence but do not mtersect the excavation.

(b) Competent rock, rock stress problems

SRF

H. Low stress, near surface >200 >13 2.5


J. Medium stress 200-10 13--D.66 1.0
K. High stress, very tight structure
(usually favourable to stability,
may be unfavourable for wall
stability) 10-5 0.66--D.33 0.5-2
L. Mild ro.ckburst (massive rock) 5-2.5 0.33--D.16 5-10
M. Heavy rockburst (massive rock) <2.5 <0.16 10-20

Notes: (Ii) For strongly anisotropic virgin stress field (If measured): when
5~0't!0'3~ 10, reduce O'er and O'Tf to 0.80'er and 0.80'Tf. When O'ti0'3> 10,
reduce O'er and O'Tf to 0.60'e and 0.60'Tf> where O'er = unconfined compression
strength, and O'Tf= tensile strength (point load) and 0' 1 and 0' 3 are the major
and minor principal stresses.
(III) Few case records available where depth of crown below surface is less than
span width. Suggest SRF mcrease from 2.5 to 5 for such cases (see H).

(c) Squeezing rock; plastic flow of incompetent rock under the influence
of high rock pressure

N. Mild squeezing rock pressure 5-10


O. Heavy squeezing rock pressure 10-20

182
Behaviour of Rock Masses 183

Table 7.4-continued

ld) Swelllng rock; chemical swelling actIvity depending on presence of


water

P. Mild swelling rock pressure 5-10


R. Heavy swelling rock pressure 10-15

Adduional notes,'
When makmg estimates ofthe rock mass quahty (Q), the followmg gUldehnes should be
followed in addition to the notes hsted under parts 1 to 6:
(1) When borecore is unaVailable, RQD can be estimated from the number ofjomts
per unit volume, in which the number of joints per metre for eachJoint set are added. A
sImple relatIOn can be used to convert this number to RQD for the case of clay-free rock
masses:
RQD= 115-3.3Jv
where Jv is the total number of joints per cubic metre (RQD = 100 for Jv < 4.5).
(2) The parameter I n representing the number of joint sets will often be affected by
foliatIOn, schistocity, slatey cleavage or bedding, etc. If strongly developed, these
parallel 'Joints' should obviously be counted as a complete joint set. However, if there
are few 'jomts' visible, or only occasIOnal breaks m borecore due to these features, then
It will be more appropriate to count them as 'random joints' when evaluating I n in
part 2.
(3) The parameters J, and J. (representing shear strength) should be relevant to the
weakest significant joint set or clay:filled discontinuity m the gIven zone. However, If the
joint set or discontinuity wIth the minimum value of (J,/J.) is favourably oriented for
stability, then a second, less favourably orientated ]omt set or discontinuity may
sometimes be of more significance, and its higher value of J,/J. should be used when
evaluating Q. The value of J,/J. should in fact relate to the surface most lIkely to allow
failure to imtiate.
(4) When a rock mass contains clay, the factor SRF appropriate to loosening loads
should be evaluated (part 6(a)). In such cases the strength of the intact rock is ofhttle
interest. However, when Jointing is mimmal and clay is completely absent, the strength
of the intact rock may become the weakest link, and the stability will then depend on
the ratio of rock stress/rock strength (part 6(b)). A strongly anisotropic stress field is
unfavourable for stabIlity and is roughly accounted for as in note (ii), part 6(b).
(5) The compressive and tensile strengths (O'c and O'T) of the intact rock should be
evaluated m the saturated conditIOn If this is appropnate to present or future in situ
conditions. A very conservative estimate of strength should be made for those rocks
that deteriorate when exposed to mOIst or saturated conditIOns.

shea~ zones or clay-bearing rock, the squeezing load in incompetent,


plastIc rocks and the rock stress in competent rocks.
It is perhaps surprising in view of the difference in approach that
there is any relation between RSR, RMR and Q, but Bieniawski
(1976) illustrates a remarkably close correlation (Fig. 7.5) between
the latter two. The use of the Q factor is principally to determine
tunnel support and, although beyond the scope ofthe present text, it is
worth illustrating. Barton et al. (1974) propose an inverse safety
184 Engineering Behaviour of Rocks

EXCEPT IQN.6.ll Y
POOR
EX',.'6i~ElY ;;M~ POOR FAIR GOOO I ~E~!
GOOD
EXI1W'ElY
GooO
exc
GOOD

./
100 1/ ~
/ ~
90
KEY
. /
/
l( //
g<>
0

. x.·/
0
"'SO N G I ASE
C STUDIES
:>: GEOM CASE STUDI ES /
0
~
'"
l:J 70
OTHER CASE STUDIES x
, ,( o 0 / '"
0

v.:;-i '"
0
V

.
:z 0
;:: I/o ,.

~o/ . :/
« 0<
</ 0 /
cr60
z • I/o /
o / .- ,.;;
."

~ so 0

w
u:: ~'\./ /. 0
o
",'- /
o x. '\ /
0 ~ f<...-. 90% CONFIDENCE
.0
Vi 40 LIMIT
VI
« O·~,<'().
<'/. • O· 0
<,/,o
d 30 ~'
q..~ ~8..« ~
~.o
~~
~".- 8
~
~~:( ~~
0

~+I/
• •
"
:'l
// ft //
0
20
:I: <

..'"
~
:>: ~
8 10
V )/ 0
l:J 0
A/ "
0.001 0.1 10 40 100 400 1000
ROCK MASS nUAlI TY a

Figure 7.5 Correlation between the Geomechanics (RMR) and Rock Mass
Quality (Q) classifications based on various case studies (after Barton 1976b).
Note the broad classifications suggested by Barton et al. (1974) for logarithmic
divisions of Q values.

factor known as the excavation support ratio which varies from 3 to 5


for temporary mine openings to 0.8 for sensitive structures such as
underground nuclear power stations, and underground structures
used for various public purposes. This is then divided into the
excavation span to give an equivalent dimension, De' in metres. Fig. 7.6
plots De against Qto give an indication of the maximum unsupported
span. Information is also given on the degree of support required for
various DJQ ratios.
The data in Table 7.4 are less important in determining with
detailed accuracy an absolute Q value than in summarizing for an
engineer who attempts to calculate a value of Q the more important
factors affecting the behaviour of rock masses. As such, Table 7.4 is an
important qualitative checklist which gives valuable information on
rock mass behaviour as well as number by which it can be classified.

7.3 Rock mass strength criterion

Rock mass classifications are only useful - except insofar as they help
Behaviour of Rock Masses 185

./
EXC EPT IONALL Y
Exr.'6~~m ~=!'.! El<I!WOElY EXC
100
POOR ~~ POOR FAIR GOOO GOOD GOOD GOOD
I 1/ • ~
/ ~
90
KEY
. /
/
v. //
<>
g
0

.
0
N G ICASE STUDIES
"'60
>:
'"
GEOM CASE STUDIES 0
~ x 0./ //

"
OTHER CASE STUDI E S x
~ 70
o • '"
0
0

. .
V
;::
i /. a. </ ~i, /
0

« 0
"'60
z
o /
6'. ) • :/
/.
0
/
/
,.
."

~ SO
~'\./ / . •o • ~....j V
ii
LJ
u:::
Vi 40
.~
.X . '\.~V f<..-. 90%LIMITFlOENCE
0 (0

'"
«
()
('/.
q.. OJ''
().
,(' ••
• ().'
('/'O
0

c:: 30 0:>'
~/ ~g.,<~ .~...
8
tJ "
~ ~~. ~~ ~+/
// ft //
20
~ '"
>:
8 10 .."'
~

'V // 0
l:I
/ 0

A "
0.001 0.01 0.1 10 40 100 400 1000
ASS (lUALI TY a ROCK
Figure 7.6 Relation between equivalent dimension and Q with a proposed
envelope for tunnel support (after Barton et al. 1974).

in the understanding of factors assisting rock deformation - if they


can be related to some characteristic of the rock mass. Thus empirical
relations between RQD or fracture index and the ratio between
laboratory and field modulus (Table 1.8) are particularly useful in
foundation design. Similarly, relations between rock mass classifi-
cations and support loads or stand-up time are useful as a starting
point in shallow underground tunnel or mine design.
In deeper mines, it is better to consider the intact strength of the
rock and to base design on the relation between this and stress, taking
into account any resistance resulting from the residual strength of
fractured rock.
In the case of slopes or large or shallow underground structures,
design can be based on an analysis of individual discontinuity shear
resistance. In some cases, however, the absence of major structural
features and the presence of strong discontinuity patterns may create
the conditions where the rock mass can be treated as a homogen-
eously fissured continuum. Such cases are probably rare, but Hoek
and Brown (1980a,b) consider in some detail the case of a rock mass
as an isotropic assembly of interlocking angular particles and develop
from this an empirical rock mass strength criterion which is worth
examining as an extension of the rock mass classification systems.
186 Engineering Behaviour of Rocks

Hoek and Brown use as their basic strength envelope a curve


similar to that m equation (3.14) for a Griffith crack in plane
compreSSIOn:

(7.8)

where I1cf is the uniaxial compressive strength of the intact rock


material and m, S are constants which depend on the properties of the
rock and on the extent to which it has been disturbed or broken before
being subjected to failure stresses 11 If and 113'
Then by putting 11 3= 0 in equation (7.8), the uniaxial compressive
strength of the rock mass is given by:
I1cm= I1cfJS (7.9)
and by putting 11 If = 0, the uniaxial tensile strength of the rock mass is
given by:
I1Tm =tl1cf[m-J(m 2 +4S)] (7.10)
These two equations define the limiting values of S. Thus if S = 1,
11 cm = 11 cf'
the value for intact rock. If S = 0, 11cm = 11Tm = 0, as would be
expected for completely broken rock. Thus for intermediate stages
between intact and broken rock, S will have values between 1 and O.
The constant m may be considered as an empirical curve-fitting
parameter affecting the slope of the curve, and by putting S = 1 it is
possible to calculate values from 5 to 7 for limestones, dolomites and
mudstones, to 23 to 28 for coarse-grained igneous rocks. A full range
of values related to Bieniawski's (1973) and Barton et al.'s (1974)
classifications is given in Table 7.5. This suggests a very powerful
analytical tool for certain rock conditions. It is, however, very
important to define these conditions. Most simply the conditions
which are excluded are where major structural features control
stability, or where the rock in a broken or intact state does not
behave in a wholly brittle manner.
In Chapter 4 the concept of brittle-ductile transition has been
considered, together with Barton's (1976a) suggestion that broken or
intact rocks behave in a ductile manner when 11 If = 211 3 (saturated
rock) and I1lf';;;3113 (dry, drained rock). Hoek and Brown quote Mogi's
(1966b) brittle-ductile transition line of 11 1 = 3.411 3' If this is plotted
on the recommended envelopes for limestone, dolomite and marble
(Fig. 7.7) it can be seen that only in the intact state does this material
behave in a brittle manner at high confining pressures and tbat in
Table 7.5 m and S values, after Hoek and Brown (1980b)

Dolomite, Mudstone, Andesite, Gahhr()~


Rock Approx. Approx. limestone, siltstone, Sandstone dolerite, gneiss,
quality RMR* Qt marble shale, slate quartzite rhyolite granite

m S m S m S m S m S

Intact
Laboratory sample 100 500 7 10 15 17 25 II

Very good
Interlocked, undisturbed, 85 100 3.5 0.1 5 0.1 7.5 0.1 8.5 0.1 12.5 0.1
unweathered joints at 3 m

Good
Slightly weathered and 65 10 0.7 0.004 0.004 1.5 0.004 1.7 0.004 2.5 C).004
disturbed joints 1-3 m

Fair
Several sets 44 0.14 10- 4 0.2 10- 4 0.3 10- 4 0.34 10- 4 0.5 110- 4
moderately weathered
joints at 0.3-1 m

Poor
Numerous weathered 23 0.1 0.04 10- 6 0.05 10- 6 0.08 10-" 0.09 10- 6 0.13 10- "
joints at 30-50 mm

Very poor
Numerous heavily 0.01 0.007 0 om 0 oms 0 0.017 0 0.025 0
weathered joints
at 50 mm

(1973) Rock Mass Rating.


* Bieniawski's
t NGI Rock Mass Quality (Barton el al. 1974).
4

2
C1 f
ere

S = 10 - , Fair

S .10- 6 Poor
S =0 Ver poor

Figure 7.7 Datafor dolomite, limestone and marble,for the rock qualities in
Table 7.5 plotted in terms of equation (7.S). Note that using Mages (1966b)
criterion for ductility, 0'1 = 3.40' 3, only intact and very good rock behave in a
brittle manner.
Behaviour of Rock Masses 189

anything less than the 'very good' state it behaves throughout its
range in a ductile manner.

7.4 The relevance of rock mass strength

Discussion of Hoek and Brown's rock mass strength criterion leads


to the general question which recurs constantly in rock eng-
ineering - which is the most important type of behaviour, that of
intact rock or of massive rock? It is for instance normally assumed
that in rock masses the deformation and strength of a rock are
controlled by the structural discontinuities which exist in massive
rock. As illustrated in Chapter 1 and particularly Tables 1.6 and 1.7,
Terzaghi, who was more observant than most, showed that loosening
above tunnels in discontinuous rocks caused larger support loads than
occurred in intact rocks. In addition, work on pillars leading to the
size-strength relations of equation (1.19) and Fig. 1.12 has also
indicated that the strength of rocks in pillars tends to reduce with size.
The implication is that this results from the increased relevance of
discontinuities with increasing size. It is significant that Hoek and
Brown (1980a) use the example of mine pillars as a particular
illustration of their rock mass strength criterion. It can, however, be
argued that neither tunnel roofs nor pillars are good examples of
underground structures to treat as a general case. For instance, blocks
of rock in tunnel roofs are subjected to gravity and changes in
environmental conditions - particularly as a result of water flowing
to a new low potential surface. It is not surprising that the properties
of the discontinuities change. Pillars are similarly subjected to
weathering, but in addition they are unique among rock structures in
being uniaxially loaded, and low width/height ratio pillars are
virtually the only underground examples of 'soft' loading.
It is worth examining the justification for the strength-size
equation, confirmed by Bieniawski (1981) in the form:

(7.11)
for various rocks.
An equation similar to this can be derived (see Tsur-Lavie and
Denekemp 1982) by considering energy released from a cube during
190 Engineering Behaviour of Rocks

prior to fracture will be given by:

(7.12)

where E is the deformation modulus.


Then if at fracture an amount of energy Up is required to satisfy
each unit area offracture surface created, and for a given material the
fracture surface resulting from brittle breakdown is assumed to have
an area proportional to the cross-sectional area (L2) of the cube,

Uf a;fL
Up = L2 E (7.13 )

and if Up is constant and independent of size for a given rock, then


acf = kL - 0.5 which is the form of equation (7.11).
However, if the material behaves in an essentially ductile manner,
then instead of fractures being initiated at failure, energy will be
evenly distributed throughout the volume of the material, causing
homogeneous plastic distortion. In that case equation (7.13) can be
written:

(7.14)

and if Uv is constant and independent of size for a given rock, then

(7.15)
and there will be no size effect.
This very simple analysis is useful in that it indicates that, rather
than being a feature of discontinuities in the specimen, size-strength
effects in pillars are related to the method of breakdown of the pillar,
which may in turn be related to the shape of the pillar and to the
method of testing in the case of laborat )fy-scale pillars.
The implication of this very simple analysis is that discontinuities
are only important in rock engineering where they constitute a
potential failure plane or zone of weakness. The majority of
discontinuities so described would be either faults or shear zones. In
all other cases it can be argued that short-term rock behaviour will be
controlled by the performance of the intact rock when subjected to
the stresses redistributed by excavation. The long-term behaviour at
surfaces and particularly excavation roofs, overh;mgs and sidewalls
will of course be affected by loosening under the effects of gravity and
Behaviour of Rock Masses 191

changing environment, and this must be taken into account in


support and structural design. However, except insofar as they
predict and help to understand these long-term features, criteria for
the description of rock mass properties may lead to confusion and
overdesign, rather than improved design in rock.
References

Atkinson, J. H. and Bransby, P. L. (1978) The Mechanics of Soils; An


Introduction to Critical State Soil Mechanics, McGraw-Hill, London.
Attewell, P. B. and Farmer, I. W. (1973) Fatigue behaviour of rocks. Int. J.
Rock. Mech. Min. Sci., 10, 1-9.
Attewell, P. B. and Farmer, I. W. (1976) Principles of Engineering Geology,
Chapman and Hall, London.
Attewell, P. B. and Sandford, M. R. (1974) Intrinsic shear strength of a brittle
anisotropic rock. Int. J. Rock. Mech. Min. Sci., 11,423-51.
Badger, C. W., Cummings, A. D. and Whitmore, R. L. (1956) The
disintegration of shales in water. J. Inst. Fuel, 29, 417-23.
Balla, A. (1960) Stress conditions in triaxial compression. J. Soil Mech.
Found. Div., A.S.C.E., 86, (SM6), 57-84.
Barton, N. (1973) Review of a new shear strength criterion for rock joints.
Eng. Geol., 7, 287-332.
Barton, N. (1976a) The shear strength of rock and rock joints. Int. J. Rock
Mech. Min. Sci., 13,255-79.
Barton, N. (1976b) Recent experiences with the Q-system of tunnel support
design. Proc. Symp. Expl. Rock Eng., Johannesburg, Balkema, Cape Town,
Vol. 1, pp. 107-18.
Barton, N., Lien, R., and Lunde, J. (1974) Engineering classification of rock
masses for the design of tunnel support. Rock Mech., 6, 189-236.
Bieniawski, Z. T. (1967) Mechanism of brittle fracture of rock, Parts 1,2 and
3. Int. J. Rock Mech. Min. Sci., 4, 395-430.
Bieniawski, Z. T. (1970) Time dependent behaviour of fractured rock. Rock
Mech.,7, 123-37.
Bieniawski, Z. T. (1973) Engineering classification of jointed rock masses.
Trans. S. Afr. Int. Civ. Eng., 15, 335-44.
Bieniawski, Z. T. (1974) Estimating the strength of rock materials. J. S. Afr.
Inst. Min. Metall., 74, 312-20.
Bieniawski, Z. T. (1976) Rock mass classifications in rock engineering. Proc.
Symp. Expl. Rock Eng., Johannesburg, Balkema, Cape Town, Vol. 1, pp.
97-106.

193
194 Engineering Behaviour of Rocks

Bieniawski, Z. T. (1981) Improved design of coal pillars for US mining


conditions. Proc.1 st Ann. Con! Ground Control in Mining, W. Virginia, pp.
12-22.
Bishop, A. W. (1959) The Principles of Effective Stpm, Tek. UkeblRd, No. J9,
Oslo.
Bishop, A. W. (1972) Shear strength parameters for undisturbed and
remoulded soil specimens. Proc. Symp. Stress-Strain Behaviour of Soils,
Oxford, Foulis, Yeovil, pp. 3-58.
Boughton, N. O. (1968) Correlation of measured foundation modulus with in
situ rock properties. Proc. Int. Symp. Rock Mech., Madrid.
Brace, W. F. (1960) An extension ofthe Griffith theory offracture to rocks. J.
Geophys. Res., 65, 3477-80.
Brace, W. F., Paulding, B. W. and Scholz, C. S. (1966) Dilatancy in the
fracture of crystalline rocks. J. Geophys. Res., 71, 3939-53.
Bridgeman, P. W. (1928) The compressibility ofthirteen natural crystals. Am.
J. Sci., 15, 287-96.
Bridgeman, P. W. (1949) Volume changes in the plastic stages of simple
compression. J. Appl. Phys., 20, 1241-51.
Broch, E. and Franklin, 1. A. (1972) The point load strength test. Int. J. Rock
Mech. Min. Sci., 9, 669-97.
Brook, N. (1977) A method of overcoming both size and shape effects in point
load testing. Proc. Con! Rock Eng., British Geotechnical Society,
Newcastle upon Tyne, pp. 53-70.
Brown, E. T. (1981) (ed.) Rock Characterisation, Testing and Monitoring,
Pergamon, Oxford.
Brown, E. T. and Hoek, E. (1978) Trends in relationships between measured
in situ stresses and depth. Int. J. Rock Mech. Min. Sci. Geomech. Abstr., 15,
211-15.
Brown, E. T. and Michelis, P. (1978) A critical state yield equation for strain
softening rock. Proc.19th US Rock Mech. Symp., Lake Tahoe, pp. 515-19.
Carter, P. and Sneddon, M. (1977) Comparison of Schmidt hammer, point
load and unconfined compression tests in Carboniferous strata. Rock
Engineering, British Geotechnical Society, Newcastle-upon-Tyne, pp.
197-210.
Cecil, O. S. (1970) Correlation of rock bolt-shotcrete support and rock
quality parameters in Scandinavian tunnels. Proc. Swedish Geotech. Inst.,
Stockholm, No. 27.
Coates, D. F. (1964) Classification of rocks for rock mechanics. Int. J. Rock
Mech. Min. Sci., 1,421-31.
Coates, D. F. (1965) Rock Mechanics Principles, Mines Branch Monograph,
874, Canadian Department of Energy, Mines and Resources, Ottawa.
Cook, N. G. W. (1965) The failure of rock. Int. J. Rock Mech. Min. Sci., 2,
389-403.
Coon, R. F. and Merritt, A. H. (1970) Predicting in situ modulus of
deformation using rock quality indexes. In situ Testing of Rock, ASTM,
STP477, pp. 154-73.
Crouch, S. L. (1970) The Influence oj Failed Rock on the Mechanical
Behaviour oj Underground Excavations. Ph.D. Thesis, Univ. of Minnesota.
References 195

Cruden, D. M. (1971) The form of the creep law for rock under uniaxial
compression. Int. J. Rock Mech. Min. Sci., 8, 105-26.
D'Andrea, D. A., Fischer, B. L. and Fogelson, D. E. (1965) Prediction of
Compressive Strength from Other Rock Properties, US Bur. Mines, Rep.
Invest. 6702.
Daw, G. P. (1971) A modified Hoek-Franklin triaxial cell for rock
permeability measurements. Geotechnique, 21, 89-91.
Deere, D. U. (1964) Technical description of cores for engineering purposes.
Rock Mech. Eng. Geol., 1, 16-22.
Deere, D. U. (1966) Contribution to discussion. Proc.1st Congo Int. Soc. Rock
Mech., Lisbon, Vol. 3, pp. 156-58.
Deere, D. U. (1968) Geotechnical considerations, in Rock Mechanics in
Engineering Practice (eds. K. G. Stagg and o. C. Zienkiewcz), Wiley,
London. '
Deere, D. u., Hendron, A. J., Patton, F. D. and Cording, E. J. (1966) Design of
surface and near surface construction in rock. Proc. 8th US Symp. Rock
Mech., Minneapolis, pp. 237-303.
Deere, D. U. and Miller, R. P. (1966) Engineering Classification and Index
Properties for Intact Rock. Air Force Weapons Lab. Rep. AFWL-TR-65-
16, Kirkland, New Mexico.
Donath, F. A. (1964) Strength variation and deformation behaviour in
anisotropic rock. State of Stress in Earth's Crust, Elsevier, New York,
pp.281-98.
Dreyer, W. (1973) The Science of Rock Mechanics - The Strength Properties
of Rock, 2nd edn, Trans. Tech. Pubis., Clausthal.
Duncan, 1. M. and Goodman, R. E. (1968) Finite Element Analyses of Slopes
in Jointed Rock, US Army Corps of Engineers, Rep. S63-3.
Edmond, J. M. and Patterson, M. S. (1972) Volume changes during the
deformation of rocks at high pressures. Int. J. Rock Mech. Min. Sci., 9,
161-82.
Einstein, H. H. and Hirschfield, R. C. (1973) Model studies on mechanics of
jointed rock. J. Soil Mech. Found. Div., A.S.C.E., 99, 833-48.
Evans, I. and Pomeroy, C. D. (1966) Strength Fracture and Workability of
Coal, National Coal Board, London.
Farmer, I. w. (1980) Face and Roadway Stability in Underground Coal Mines:
Geotechnical Criteria, Report EUR 7298 EN, Commission the European
Community, Luxembourg.
Farmer, I. W. and Gilbert, M. 1. (1981) Time dependent strength reduction of
rock salt. Proc. 1st Conf. on Mechanical Behaviour of Rock Salt,
Pennsylvania State Univ.
Fine, 1., Tijane, S. M., Vouille, G. and Boucly, F. (1979) Determination
experiment ale de quelques parametres elastoviscoplastiques des roches.
Proc. 4th Int. Congr. Rock Mechanics, Montreux, Vol. 1, pp. 139-43.
Fookes, P. G. (1965) Orientation of fissures in stitT overconsolidated clay of
the Siwalik system. Geotechnique, 15, 195-206.
Franklin, J. A., Broch, E. and Walton, G. (1971) Logging the mechanical
character of rock. Trans. Inst. Min. Metall., 80A, 1-9.
Gerogiannopoulos, N. and Brown, E. T. (1978) The critical state concept
applied to rock. Int. J. Rock Mech. Min. Sci Geomech. Abstr., 15, 1-10.
196 Engineering Behaviour of Rocks

Ghaboussi, 1., Wilson, E. L. and Isenberg, 1. (1973) Finite element for rock
joints and interfaces. J. Soil Mech. Found. Div., A.S.C.E., 99, 833-48.
Gilbert, M. 1. (1981) Shafts in Squeezing Rock, Ph.D. Thesis, Univ. of
Newcastle-upon-Tyne.
Gilbert, M. 1. and Farmer, I. W. (1981) A time dependent model for lining
pressure based on strength concepts. Proc. Symp. Weak Rock, Tokyo, Vol.
1, Balkema, Rotterdam, pp. 137-42.
Goodman, R. E. (1974) The mechamcal properties ofjomts. Proc. 3rd Congr.
Int. Soc. Rock Mech., Denver, Vol.lA, pp. 127-40.
Goodman, R. E. (1976) Methods of Geological Engineering in Discontinuous
Rocks, West Pub!. Co., St Pau!'
Gowd, T. N. and Rummel, F. (1980) Effect of confining pressure on the
fracture behavIOur of porous rock. Int. J. Rock Mech. Min. Sci. Geomech.
Abstr., 17, 225--9.
Griffith, A. A. (1921) The phenomena of rupture and flow in solids. Phil.
Trans. R. Soc., A221, 163-98.
Griffith, A. A. (1924) Theory of rupture. Proc. 1st Int. Congr. Appl. Mech.,
Delft, pp. -55-{iJ
Griggs, D. (1939) Creep of rocks. J. Geol., 47, 225-51.
Haimson, B. C. (1978) The hydrofracturing stress measuring method and
recent field results. Int. J. Rock Mech. Min. Sci. Geomech. Abstr., 15,
167-78.
Haimson, B. C. and Kim, R. Y. (1971) Mechanical behavior of rock under
cyclic faIlure. Proc. 13th US Rock Mech. Symp., Umv. IllInois, Urbana, pp.
845-{i2.
Hanna, T. H. (1973) Foundatwn Instrumentation, Trans. Tech. Pubis.,
Cleveland. .
Hardy, H. R., Kim, R. Y., Stefanko, R. and Wang, Y. 1. (1969) Creep and
microseismic activity in geologic materials. Proc. 11th US Rock Mech.
Symp., Berkeley, pp. 377-413.
Hassani, F. P. and Scobie, M. 1. (1981) Properties of weak rocks with special
reference to the shear strength of their discontinuities as encountered in
British surface coal mining. Proc. Symp. Weak Rock, Tokyo, Vol. 1,
Balkema, Rotterdam, pp. 355-64.
Hawkes, I. and Mellor, M. (1970) Uniaxial testing in rock mechanics
laboratories. Eng. Geol., 4, 177-285.
Hawkes, I., Mellor, M. and Gariepy, S. (1973) Deformation of rocks under
uniaxial tension. Int. J. Rock Mech. Min. Sci., 10,493-507.
Hill, R. (1950) The Mathematical Theory of Plasticity, Clarendon Press,
Oxford.
Hobbs, N. B. (1973) Effects ofnon-lineanty on the prediction of settlements of
foundatIOns on rock. Q. J. Eng. Geol., 6, 153-{i8.
Hobbs, N. B. (1974) The prediction of settlement of structures on rock. Con!
Settlement of Structures, Cambridge, pp. 579-610.
Hoek, E. (1966) Rock mechanics - an introduction for the practical engineer.
Min. Mag., 114, 236-43.
Hoek, E. (1970) Estimating the stability of excavated slopes in opencast
mines. Trans. Inst. Min. Metall., 79A, 109-32.
References 197

Hoek, E. and Franklin, 1. A. (1968) Simple triaxial cell for field or laboratory
testing of rock. Trans. Inst. Min. Metall., 77A, 22--4.
Hoek, E. and Bray,J. W. (1974) Rock Slope Engineering, Institution of Mining
and Metallurgy, London.
Hoek, E. and Brown, E. T. (1980a) Underground Excavations in Rock,
Institution of Mining and Metallurgy, London.
Hoek, E. and Brown, E. T. (1980b) Empirical strength criterion for rock
masses. J. Geotech. Eng. Div., A.S.C.E., 106, 1013-35.-
Hough, B. K. (1957) Basic Soils Engineering, Ronald Press, New York.
Hudson, 1. A., Brown, E. T. and Fairhurst, C. (1971) Shape of the complete
stress-strain curve for rock. Proc. 13th US Rock Mech. Symp., Univ. of
Illinois, Urbana, pp. 773-95.
Hudson, 1. A., Crouch, S. L. and Fairhurst, C. (1972) Soft, stilT and
servocontrolled testing machines; a review with reference to rock failure.
Eng. Geol., 6, 155-89.
Jaeger, 1. C. (1969) Elasticity Fracture and Flow with Engineering and
Geological Applications, 3rd edn, Chapman and Hall, London.
Jaeger, 1. C. and Cook, N. G. W. (1969) Fundamentals of Rock Mechanics,
Chapman and Hall, London.
Jahns, H. (1966) Messung der Gebirgsfestigkeit in situ bei wachsendern
Masstabsverhiiltnis. Proc. 1st Congo Int. Soc. Rock Mech., Lisbon, Vol. 1,
pp.477-82.
Jamison, D. B. and Cook, N. G. W. (1979) An analysis of the measured values
for the state of stress in the earth's crust. J. Geophys. Res., unpublished
paper quoted in Jaeger and Cook (1969).
John, K. W. (1974) Geologists and civil engineers in the design of rock
foundations of dams. Proc. 2nd Int. Congr. Int. Assoc. Eng. Geol., Sao
Paulo, Paper VI-PC-3.
Judd, W. R. (1964) Rock stress, rock mechanics and research, in State of Stress
in the Earth's Crust (ed. W. R. Judd), Elsevier, New York, pp. 5-53.
Judd, W. R. and Huber, C. (1962) Correlation of rock properties by statistical
methods. Int. Symp. Min. Res., Pergamon, Oxford, pp. 621--48.
Krumbein, W. C. (1942) Physical and chemical changes in sediments after
deposition. J. Sed. Petr., 12, 111-17.
Ladanyi, B. and Archambault, G. (1969) Simulation of shear behaviour of a
jointed rock mass. Proc.11th US Rock Mech. Symp., Berkeley, pp. 105-25.
Ladanyi B. (1974) Use of the long term strength concept in the determination
of ground pressure on tunnel linings. Proc. 3rd Congr. Int. Soc. Rock M ech.,
Denver, Vol. I(B), pp. 1150-6.
Lama, R. D. and Vutukuri, V. S. (1978) Handbook on Mechanical Properties
of Rock-Testing Techniques and Results, Vol. III, Trans. Tech. Pubis.,
Clausthal.
Lambe, T. W. (1951) Soil Testingfor Engineers, Wiley, New York.
Lambe, T. W. and Whitman, R. V. (1979) Soil Mechanics (SI version), Wiley,
New York.
Lauffer, H. (1958) Gebirgsklassifizierung fur den Stollenbau. Geologie und
Bauwesen, 24, 46-51.
Leeman, E. R. (1964) The measurement of stress in rock, Parts I and II. J. S.
Afr. Inst. Min. Metall., 65, 45-114, 254-84.
198 Engineering Behaviour of Rocks

Leeman, E. R. and Hayes, D. 1. (1966) A technique for determining the


complete state of stress in rock using a single borehole. Proc.1 st Congr. Int.
Soc. Rock Mech., Lisbon, Vol. 2, pp. 17-24.
Londe, P. (1973) The role of rock mechanics in the reconnaissance of rock
foundations. Q. J. Eng. Geol., 6, 58-74.
McLintock, F. A. and Walsh, 1. B. (1963) Friction on Griffith cracks in rock
under pressure. Proc. 4th US Congr. App!. Mech., Berkeley, pp. 1015-21.
Mirza, U. A. (1978) Investigation Into the Design Criteria for Underground
Openings in Rocks Which Exhibit Rheological Behaviour, Ph.D. Thesis,
Univ. of Newcastle-upon-Tyne.
Mogi, K. (1966a) Some precise measurements of fracture strengths ofrocks
under uniform compressive stress. Rock Mech. Eng. Geol., 4, 41-5.
Mogi, K. (1966b) Pressure dependence of rock strength and transition from
brittle fracture to ductile flow. Bull. Earthquake Res. Inst. Tokyo Univ.,44,
215-32.
Moye, D. G. (1955) Engineering geology for the Snowy Mountain Scheme in
Australia, J. Inst. Eng. Aust., 9, 95-112.
Murrell, S. A. F. (1963) A criterion for brittle fracture of rocks under triaxial
stress. Proc. 5th US Rock Mech. Symp., Univ. of Minnesota, Pergamon,
Oxford, pp. 563-77.
Obert, L., Windes, S. L. and Duvall, W. I. (1946) Standardised Tests for
Determining the Physical Properties of Mine Rocks, US Bureau of Mines.
Report of Investigation No. 3891.
Obert, L. and Duvall, W. I. (1957) Microseismic Method of Determining the
Stability of Underground Openings, US Bureau of Mines Bull. 573.
Obert, L. and Duvall, W. I. (1967) Rock Mechanics and the Design of
Structures in Rock, Wiley, New York.
Olsson, W. A. and Peng, S. S. (1976) Microcrack nucleation in marble. Int. J.
Rock Mech. Min. Sci., 13, 53-9.
Onodera, T. F. (1963) Dynamic investigation of foundation rocks in situ.
Proc. 5th US Rock Mech. Symp., Univ. of Minnesota, pp. 517-33.
Orowan, E. (1949) Fracture and strength of solids. Repts. on Prog. in Physics,
12, 185-232.
Patterson, M. S. (1978) Experimental Rock Testing - the Brittle field,
Springer-Verlag, Berlin.
Patton, F. D. (1966) Multiple modes of shear failure in rock. Proc. 1st Congr.
Int. Soc. Rock Mech., Lisbon, Vol. 1, pp. 509-13.
Patton, F. D. and Deere, D. U. (1970) Significant geologic factors in slope
stability. Proc. Symp. Planning Open Pit Mines, Balkema, Cape Town,
pp.143-51.
Peng, S. S. and Podnieks, E. R. (1972) Relaxation and the behaviour offailed
rock. Int. J. Rock Mech. Min. Sci., 9, 699-712.
Phillips, F. C. (1971) The Use of Stereographic Projections in Structural
Geology, 3rd edn, Arnold, London.
Piteau, D. R. (1970) Geological factors significant to the stability of slopes cut
in rock. Proc. Symp. Open Pit Mines, Balkema, Cape Town, pp. 33-53.
Poole, R. W. and Farmer, I. W. (1980) Consistency and repeatability of
Schmidt hammer rebound data during testing. Int. J. Rock M echo Min. Sci.,
18,229-34.
References 199

Pratt, H. R., Black, A. D., Brown, W. S. and Brace, W. F. (1972) The effect of
specimen size on the mechanical properties of unjointed diorite. Int. J.
Rock Mech. Min. Sci., 9, 513~29.
Price, A. M. (1979) The Effect of Confining Pressure on the Post- Yield
Deformation Characteristics of Rocks, Ph.D. Thesis, University of
Newcastle-upon-Tyne.
Price, A. M. and Farmer, 1. W. (1979) Application of yield models to rock. Int.
J. Rock Mech. Min. Sci., 16, 157~9.
Price, A. M. and Farmer, 1. W. (1980) A general failure criterion for rocks.
Proc. 21st US Rock Mech. Symp., Rolla, pp. 256--{)3.
Price, A. M. and Farmer, 1. W. (1981) The Hvorslev surface in rock
deformation. Int. J. Rock Mech. Min. Sci., 18, 229~34.
Price, N. 1. (1966) Fault and Joint Development in Brittle and Semi-Brittle
Rock, Pergamon, Oxford.
Priest, S. D. and Hudson, 1. A. (1976) Discontinuity spacings in rock. Int. J.
Rock Mech. Min. Sci., 13, 134~53.
Ramsay, 1. G. (1967) Folding and Fracturing of Rocks, McGraw-Hill,
London.
Rankilor, P. R. (1974) A suggested field system of logging rock cores for
engineering purposes. Bull. Assoc. Eng. Geol., 11, 247~58.
Rinehart, 1. S. (1962) Effect of transient stress waves in rock. Int. Symp.
Mining Res. (ed. G. Clarke), Pergamon, Oxford, pp. 713~26.
Rispin, A. and Cooper, 1. (1972) The Mechanical Cutting of Rock Materials in
Relation to the Design and Operation of Tunnelling Machines and Rapid
Excavation Systems, Internal report, University of Newcastle-upon-Tyne.
Roberds, W. 1. and Einstein, H. H. (1978) Comprehensive model for rock
discontinuities. J. Geotech. Eng. Div., A.S.C.E., 104, 553--{)9.
Roberts, A. (1977) Geotechnology, Pergamon, Oxford.
Robertson, A. M. (1970) The interpretation of geological factors for use in
slope theory. Proc. Symp. Open Pit Mines, Balkema, Cape Town, pp. 55~ 71.
Rocha, M. (1976) Personal communication.
Roscoe, K. H., Schofield, A. N. and Wroth, C. P. (1958) On the yielding of
soils. Geotechnique, 8, 22~53.
Rowe, P. W. (1962) The stress dilatancy relation for static equilibrium of an
assembly of particles in contact. Proc. R. Soc., A264, 500~27.
Rowe, P. W. (1968) The influence of geologicalfeatures of clay deposits on the
design and performance of sand drains (summary). Proc. Inst. Civ. Eng., 39,
581.
Salamon, M. D. G. (1970) Stability, instability and design of pillar workings.
Int. J. Rock Mech. Min. Sci., 7, 613~31.
Schofield, A. and Wroth, C. P. (1968) Critical State Soil Mechanics, McGraw-
Hill, London.
Scholz, C. H. (1968a) Microfracturing and the inelastic deformation of rock
in compression. J. Geophys. Res., 73, 1417~32.
Scholz, C. H. (1968b) Experimental study of the fracturing process in brittle
rock. J. Geophys. Res., 73, 1447~54.
Skempton, A. W. (1954) The pore-pressure coefficients A and B.
Geotechnique, 4, 143~7.
200 Engineering Behaviour of Rocks

Skempton, A. W. (1961a) Effective stress in soils, concrete and rocks. Pore


Pressure and Suction in Soils, Butterworth, London, pp. 4-16.
Skempton, A. W. (1961b) Horizontal stresses in an overconsolidated Eocene
clay. Proc. 5th Int. Conf Soil Mech. Found. Eng., Paris, Vol. 1, pp. 351-87.
Snow, D. T. (1968) Rock fracture, spacings, openings and porosities. J. Soil
Mech. Found. Div., A.S.C.E., 94, 73-91.
Stini,1. (1950) Tunnel baugeologie, Springer Verlag, Vienna.
Stoney, S. M. and Dhir, R. K. (1977) Problems in measurement of anisotropy
in rocks. Rock Engineering, British Geotechnical Society, Newcastle-upon-
Tyne, pp. 137-54.
Talobre, 1. A. (1957) La Mechanique des Roches, Dunod, Paris.
Tapponier, P. and Brace, W. F. (1976) Development of stress-induced
microcracks in Westerly Granite. Int. J. Rock Mech. Min. Sci., 13, 103-12.
Taylor, D. W. (1948) Fundamentals of Soil Mechanics, Wiley, New York.
Terzaghi, K. (1925) Erdbaumechanik, Franz Deutricke, Vienna.
Terzaghi, K. (1943) Theoretical Soil Mechanics, Wiley, New York.
Terzaghi, K. (1946) Rock Tunnelling with Steel Supports, Youngstown, Ohio,
Commercial Shearing and Stamping Co.
Terzaghi, R. D. (1965) Sources of error in joint surveys. Geotechnique, 15,
287-304.
Terzaghi, K. and Richart, F. E. (1952) Stresses in rocks about cavities.
Geotechnique, 3, 57-90.
Tourenq, C, Fourmaintraux, D. and Denis, A. (1971) Propagation des ondes
et discontinuites des roches. Proc. Int. Symp. Rock Fracture, Nancy.
Tsur-Lavie, Y. and Denekamp, S. A. (1982) Size and shape effect in pillar
design. In Strata Mechanics: Developments in Geotechnical Engineering (ed.
I. W. Farmer), Vol. 32, Elsevier, Amsterdam, pp. 245-8.
Vaughan, P. R. (1963) Contribution to discussion, Grouts and Drilling Muds
in Engineering Practice, British Geotechnical Society, London, p. 54.
Ward, W. H., Burland, 1. B. and Gallois, R. W. (1968) Geotechnical
assessment of a site at Mundford, Norfolk, for a large proton accelerator.
Geotechnique, 15, 321-44.
Wawersik, W. R. (1968) Detailed Analysis of Rock Failure in Laboratory
Compression Tests, Ph.D. Thesis, Univ. of Minnesota.
Weibull, W. (1952) A survey of statistical effects in the field of material failure,
Appl. Mech. Rev., 5, 449-51.
Wickham, G. E., Tiedemann, H. and Skinner, E. H. (1972) Support
determinations based on geologic predictions. Proc. 1st Rapid Exc. Tunn.
Conf, New York, A.I.M.E., pp. 43--64.
Zismann, W. A. (1933) Compressibility and anisotropy of rocks at and near
the earth's surface. Proc. Nat. Acad. Sci., 19,666-79.
Author Index

Atkinson, 1. H., 85, 86, 90, 114 Cmden, D. M., 127


Attewell, P. B., 9, 10, 23, 48, 69,
D'Andrea, D. A., 16,21
137, 138, 142, 146, 147, 154, 156,
Daw, G. P., 22
167
Deere, D. u., 9,11,13, 15-17,28,
Archambault, G., 163
30, 31, 163, 168, 169, 171
Badger, C. W., 16 Denekamp, S. A, 189
Balla, A., 7, 60, 61, 75 Dhir, R. K., 155, 157
Barton, N., 92, 93, 108, 163, 164, Donath, D. A, 154, 156
165,172,175,178-187 Dreyer, W., 125
Bieniawski, Z. T., 16,27,28, 72, Duncan, 1. M., 26
128,129,172,174,175-178,183, Duvall, W. I., 33, 66
184, 186, 189 Edmond,1. M., 94,114
Bishop, A. W., 37, 89 Einstein, H. H., 160, 161, 165, 166
Boughton, N. 0., 30, 31 Evans, I., 27
Brace, W. F., 65, 71, 74
Bransby, P. L., 85, 86, 90, 114 Fairhurst, c., 59
Bray, 1. W., 15, 23, 146, 150, 167 Farmer, I. W., 9, 10, 16, 23, 48, 69,
Bridgeman, P. W., 38, 39, 95 97, 111, 112, 113, 116-118, 127,
Broch, E., 16 130-133, 137, 138, 142, 146, 147,
Brook, N., 14 167
Brown, E. T., 5, 6, 9, 51, 54, 55, 60, Fine, 1., 128
92, 94, 115, 144, 160, 172, 185, Fookes, P. G., 143
186, 187, 189 Franklin, 1. A., 14, 16, 94
Carter, P., 16 Gerogiannopoulos, N., 92, 115
Cecil, O. S., 175 Ghaboussi, 1., 165
Coates, D. F., 9, 33 Gilbert, M. 1., 127, 128, 129,
Cook, N. G. W., 33, 53, 56-59, 67, 130-133, 134
68, 72, 140, 158, 159 Goodman, R. E., 26, 163, 165, 167
Coon, R. F., 31 Gowd, T. N., 114
Cooper, I., 21, 22 Griffith, A A., 71
Crouch, S. L., 95 Griggs, D., 125

201
202 Author Index

Haimson, B. C, 57, 139 Peng, S. S., 74, 128


Hanna, T. H., 23 Phillips, F. C, 148, 150
Hardy, H. R., 120, 121, 122 Piteau, D. R., 145, 146
Hll~~!mi, F. P., 1& Podnieh, E. R., 118
Hawkes, I., 7, 8, 9, 60, 62 Pomeroy, CD., 27
Hayes, D. 1., 58 Poole, R W., 16
Hill, R, 87 Pratt, H. R., 28
Hirschfield, R C, 161 Price, A. M., 77, 94, 97, 99-103,
Hobbs, N. B., 11, 12, 30, 31 105-107,111,112,113,116--118,
Hoek, E., 15,23,51,52,54,55,83, 127
94, 144, 146, 150, 163, 167, 172, Price, N. 1., 10
185, 186, 187, 189 Priest, S. D., 146, 169, 170, 171
Hough, B. K., 21
Ramsay, 1. G., 43
Hudson,1. A., 4,63,64,97, 146,
Rankilor, P. R, 16
169, 170, 171
Rispin, A, 21, 22
Huber, C, 21, 141
Roberds, W. 1., 161, 165, 166
Jaeger, 1. C, 33, 53, 57-59, 72, Roberts, A., 57, 58
140, 158, 159 Robertson, A. M., 146, 147, 148
Jahns, H., 28 Rocha, M., 169
Jamison, D. B., 54, 56 Roscoe, K. H., 89
John, K. W., 145 Rowe, P. W., 3, 114
Judd, W. R., 21, 33, 141 Rummel, F., 114
Salamon, M. G. D., 64
Krumbein, W. C, 2
Sandford, M. R, 154, 156
Kim, R Y., 139
Schofield, A., 49, 86, 89, 90, 93
Ladanyi, B., 125, 135, 163 Scholz, C. H., 67, 69, 75, 76, 86, 87,
Lama, R. D., 125 88, 92, 93, 104, 108, 114, 115
Lambe, T. W., 22,119 ScobIe, M. 1., 18
Lauffer, H., 175, 178 Skempton, A W., 38-41, 54
Leeman, E. R, 57, 58 Sneddon, M., 16
Londe, P., 144, 145, 169 Snow, D. T., 23, 169
Stini, 1., 26
McLintock, F. A., 73 Stoney, S. M., 155, 157
Mellor, M.; 7,8,60,62, 74
Merritt, A. H., 31 Talobre, 1. A., 53
Michelis, P., 92 Tapponier, P., 74
Miller, R P., 9,11,13,15-17,31 Taylor, D. W., 114
Mirza, U. A., 127 Terzaghi, K., 24-28, 37, 54, 119,
Mogi, K., 75, 93, 186, 188 168,169,172,189
Moye, D. G., 26 Terzaghi, R. D., 143, 147
Murrell, S. A. F., 73 Tourenq, C, 29
Tsur-Lavie, Y., 189
Obert, L., 8, 33, 66
Olsson, W. A., 74 Vaughan, P. R., 23
Onodera, T. F., 28, 30, 168 Vutukuri, V. S., 125
Orowan, E., 71
Walsh,1. B., 73
Patterson, M. S., 94, 114, 128 Ward, W. H., 26
Patton, F. D., 162, 163 Wawersik, W. R, 65
Author Index 203

Weibull, W., 27 Wroth, P. W., 49, 86, 89, 90, 93


Whitman, R. Y., 119
Zismann, W. A., 38, 39
Wickham, G. E., 172-175
Subject Index

Angle of friction Cohesion, 7


discontinuity, 160-165 Compressibility, 38-42
internal, 7, 81-85 Compressive strength
peak, 94-111 criteria, 82, 83
residual, 94-111 discontinuity, 164-165
sliding, 151-165 magnitude, 11
Anisotropy, effect of triaxial, 7, 97-108
on elasticity, 157 uniaxial, 5, 7-14
on strength, 154-157 Consolidation, 119
Arching theory, 24-26 Coulomb criterion, 81-85
Asperities, effect of Cracks
on discontinuity cohesion, initiation, 70-74
162-164 propagation, 74-76
on discontinuity friction, Cracking, 66
162-164 Creep
Attenuation, wave energy, 142 curves, 121-124, 127
equations, 127
Biaxial strain, 46
phenomenological models,
Biaxial stress field, 44, 71, 151, 154
125-128
Bingham model, 126
primary, 124
Blasting, 142
rheological models, 125-126
Borehole stress measurement, 58
secondary, 125
Brittle behaviour, 65-80, 186, 189,
strain, 119-128
190 tertiary, 125
Brittle fracture, 69-80
Critical state
Brittleness, index, 11 concept, 89-94
Brittle-ductile transition, 86, 93, 160
definition, 85
Bulk modulus, 47, 140
discontinuity model, 165-167
Burgers model, 126
Hvorslev surface, 90-93, 111-118
Cartesian axes, 34, 35 line, 80
Coefficient of earth pressure at rest, model, 80-82
48, 53-55 Roscoe surface, 90, 91

204
Subject Index 205

stability line, 80, 109 Elasticity, 47-51


yield surface, 111, 112 Elastic waves, 140-142
Cyclic loading, 135-139 Ellipse
crack as an, 71-74
Darcy's law, 21
strain, 43
Deformation
End effects, 7, 59--63
complete curve, 77-81
Energy
description of rock, 59-80
crack surface, 71, 72
mechanics of rock, 65-80
Diagenesis, 2
during fracture, 81
release during testing, 60, 64, 65,
Dilation
189, 190
definition, 43
discontinuity, 160-166 Failure
magnitude, 188 criteria for, 81-85
mechanics of, 65-80 definition, 85
relation to microseismicity, through deformation, 85-94, 136
67-70, 120-123 through shear, 77-80, 85,
Discontinuity 98-101, 136
asperities, 162-163 Failure criteria
classification, 145 Coulomb, 81-85
cohesion, 160-165 Griffith, 71-74, 82,186
critical state model, 165-167 McLintock-Walsh, 60, 62, 73,
frequency, 169-172 74, 82
frictional resistance, 160-165 Mohr, 82-85
orientation data, 143-150 Fatigue, 137-139
porosity, 23, 29 Faults, 143, 150, 152, 190
roughness, 160-165 Field tests, 6, 14-17
shear resistance, 151, 154-165 Finite strain, 43
shear stability, 158--159 Fracture
spacing, 169-170 brittle, 65-80
stereographic projection, frequency, 31
148-153 index, 169
survey, 146-148 hydraulic, 57, 58
Distortion, 43 porosity, 29
Ductile behaviour, 85-94, 107-118, Friction
131-135, 165-167 crack, 73
Dynamic discontinuity, 151-165
modulus, 11, 140-142 internal, 7, 81-85
strength, 139 sliding, 160-165
Earthquakes, 142 Geomechanics classification, 172,
Effective stress, 37-42, 53, 151 176-178
Elastic constants Geostatic stress
bulk modulus, 47-50 coefficient of, 48, 53-55
dynamic, 11, 140-142 horizontal, 53-55
elastic modulus, 47-50 vertical, 47-53
Poisson's ratio, 47-51 Griffith cracks, 70-72
shear modulus, 47-50 Griffith criterion, 71-74, 82, 186
tests, 6, 12, 13, 65--67
Elastic behaviour, 47-51, 77, 78 Heim's rule, 53, 54
206 Subject Index

Homogeneous strain, 86-88, 190 Physical properties, 5, 18-21


Hookean material, 125 Pillars
Hvorslev surface, 90--93 brittle fracture, 135, 136,
Hydrofracture, 58 189-191
Hydrostatic stress, 41, 94 plastic failure, 135, 136, 189-191
Hysteresis, 139 Plane of weakness
effect on strength, 151-158
Igneous rocks, 1, 21
shear resistance, 158-165
Internal friction, 7, 81-85
Plastic behaviour, 85, 111-118,
Joints, 143-145 131-135, 165-166
Plastic deformation, 87, 111-118
Kelvin-Voight model, 126
Plastic flow, 87,111-118
Kozeny-Carman equation, 22
Plasticity
Laboratory tests, 6--9, 94-98, 167 associated flow rule, 114
Luders lines, 87 flow rule, 111-113
Lugeon,24 hardening law, 111-112
normality condition, 112-114
McLintock-Walsh criterion, 60, 62,
potential function, 112-114
73, 74, 82
Point load index, 14-16
Macrofracture, 74, 75
Poisson's ratio, 6,47-51,87, 88
Maxwell model, 125
Pore pressure, 37---42
Metamorphic rocks, 21
Porewater
Microcracks, 66--71
pressure, 37---42
Microseismic events, 66-71,
pressure parameters, 41, 42
120-123
Porosity, 18-22
Modulus
Prager model, 126
bulk, 47, 140
Principal stresses, 36, 37, 47
deformation, 6
dynamic, 140--142
elastic, 47 RQD,28-32, 168-172
of rigidity, 47, 140 Rheological models, 125-126
secant, 11, 14 Rock classification
strength ratio, 11-14 geomechanics, 172, 176-178
tangent, 11, 14 mass, 25-28
Modulus ratio, 11-14 NGI,178-184
Mohr circle, 45, 46, 83, 157 rock structure rating, 172-175
Mohr envelope, 83 strength, 10
Mohr criterion, 82-85 Rock deformation
NGI classification, 178-184 mechanics of, 64-80
Newtonian material, 125 Rock mass, 3
Normal stress, 33-37 behaviour, 168-191
classification, 25-28, 172-184
Octahedral stress, 60, 61 factor, 30--32, 168
Permeability strength criterion, 184-189
coefficient of, 21-24 Rock mass strength
packer test, 23 criterion, 184-189
rock mass, 22-23 relevance of, 181-191
Phase relations, 18-20, 37 Rock matrix, 41
Phenomenological models, 125-127 Rock mechanics, 1, 3
Subject Index 207

Rock quality ghear slra;n, 43


designation, RQD, 28-32, Shear strength, 11, 82, 83
168-172 Shear stress
geomechanics classification, components, 35-36
176-178 . discontinuities, 151-165
NGI classification, 178-184 representation, 44-45
Rock structure rating, 172-175 strength criteria, 81-85
Rock testing Shock wave, 140
categories, 5, 6 Size-strength relations, 27-28, 189,
compressibility, 39 190
in compression, 59-65 Slake durability test, 16-18
data, 94-111 Sliding friction, 160-165
discontinuity resistance, 167 Soil mechanics, 1
field, 14-18 Soil mineralogy, 1
in situ, 6 Specific surface energy, 71
jacketed, 38, 39 Specimens
laboratory, 5-21 description, 9
loading rates, 9 end restraint, 8, 59-63
machines, 63-65 preparation, 7, 8
rock properties, 3 shape, 7, 60
servo-controlled, 63-65 size, 7, 14
specifications, 5, 6 stress distribution in, 61, 62
stable loading, 64 Stand up time, 178
strain rates, 97, 128-131 Stereographic projection
stress measurement, 57-58 contoured data, 152, 153
triaxial cell, 94-97 discontinuity representation, 149
uniaxial,7 lower hemisphere, 150
unjacketed, 38, 39 Schmidt net, 149
Rockbursts, 142 upper hemisphere, 150
Roscoe surface, 90, 91, 115 Wulff net, 149
Rotation, 43
Stiff testing
St. Venant substance, 125 concept, 60, 63
Saturation, 20, 42 data, 97-111
Schmidt hammer, 15-18 machine design, 63-65
Schwedoff model, 126 servo systems for, 63
Sedimentary rocks Strain
composition, 1, 2 at a point, 42, 43
diagenesis, 2 axial, 66, 97-111
porosity, 21 creep, 119-128
Sediments, 1 ellipse, 43
Seismic waves finite, 43
energy, 142 hardening, 85, 160
particle velocity, 142 lateral, 66
velocity, 29, 140, 141 representation, 46
Servo-controlled test machine, shear, 43
63-65 softening, 84, 85, 160
Shear resistance time dependent, 119-142
discontinuities, 151-165 volumetric, 38, 43, 50, 51, 66,
intact rocks, 81-108 97-111
208 Subject Index
Strain ellipse, 43 tensor relations, 45--48
Strain energy, 60, 64, 65, 189 triaxial, 49
Strain gauges, 58, 65 in triaxial compression, 88
Strain hardening, 85, 160 uniaxial, 49, 50
Strain rate, 97, 128-131 Surface energy, 71
Strain softening, 80
Strength Tensile strength, 11, 82
criteria, 81-85 Terzaghi's rock classification,
effect of mineralogy, 10 23-27
effect of shape, 3, 4, 8 Testing see Rock testing
effect of size, 4, 27, 28, 189 Testing machines, 63-65
effect of strain rate, 128-131 Thrust fault, 54, 56
effect of unit weight, 21 Time dependency
envelope, 81-85, 97-111, 132 creep concepts, 124-128
failure, 81 data, 120-123
magnitudes, 11 rapid loading, 139-142
peak, 94-111 strain rate effects, 128-131
residual, 94-111 strength reduction, 131-135
rock mass strength, 184-189 Tresca criterion, 88, 89
Stress Triaxial cell, 94-97
at a point, 33-37 Triaxial stress field, 157-159
biaxial,44 Triaxial tests, 47, 97-111
cartesian tensor, 46
deviatoric, 42, 109 Uniaxial tests, 4-16
Unit weight, 20, 21
effective, 37--42, 53
engineering notation, 34
Velocity index, 28
hydrostatic, 41
Void ratio, 20
invariants, 46
Volumetric strain, 43, 65-80, 88,
Mohr circle, 45, 46 97-111, 120-123
octahedral, 60
Von Mises criterion, 89
path, 46, 48, 109-111
principal, 36, 37 Waves
principal planes, 36 body, 140
representation, 45 velocity, 29, 140, 141
sign convention, 35 seismic, 29, 140-142
tangential, 60 shock, 140
tensor, 36, 37, 46, 47 surface, 140
total, 37--42, 53
Yield
vector, 34
critical state, 86
Stress measurement, 57, 58
curves, 109-111, 132
Stress path, 46, 48, 109-111
effect of confining pressure, 86
Stress tensor, 36, 37, 45
Stress dilatency equation, 114 failure through, 85-87
plastic, 86
Stress-strain relations
Yield criteria, 85-89
complete curves, 77-80, 97-109
effect of strain rate, 128-131 Zenner model, 126

S-ar putea să vă placă și