Вы находитесь на странице: 1из 30

SHEAR STRENGTH OF ROCK JOINTS AND ROCK MASSES

Mahendra Singh
Professor, Department of Civil Engineering
IIT Roorkee, Roorkee-247 667
1. GENERAL
Assessment of shear strength of rock joints and rock masses is probably the most
important aspect in limiting equilibrium analysis of rock slopes. A small change in shear
strength may substantially affect the safe angle of slope. Whereas the prediction of shear
stress acting along a failure surface is relatively easy, it is very difficult and cumbersome
to assess the shear strength with confidence. Some of the reasons behind this are:
i)

The failure of rock mass may take place due to sliding along the existing
joints, shearing of the rock material, rotation of blocks of rock material, or due to
a complex combination of all of these modes.

ii)

Jointed rocks and rock masses may behave anisotropically due to the presence
of discontinuities, even though the rock material itself may be isotropic.

iii)

The shear strength along a failure surface depends on the level of normal
stress, and varies in a non-linear manner. This aspect is much significant for
slopes, as compared to other structures like foundations and tunnels. Non-linearity
in shear strength behaviour is very high at low normal stress levels, the condition
which generally prevails in case of slopes.

iv)

Direct shear test results, even if performed in-situ, may not be directly
applicable due to scale effect, kinematics and in-homogeneity of the slope.
This paper attempts to address some of these issues. It is assumed herein that the

stresses involved are effective stresses and the shear strength parameters applicable are
expressed in terms of effective stresses. For simplicity, the prime sign () generally used
to denote effective stress, and parameters in terms of effective stress will be dropped

throughout this chapter. Further, the stresses are considered static and not dynamic in
nature.
1.1.

Classes of Rock Shear Strength


The shear strength aspect of rocks may be grouped into the following two major

categories:
i)

Shear strength along planar discontinuity (Discontinuity shear strength)

ii)

Shear strength of jointed rock mass (Rock mass shear strength)


(a)

Isotropic rock mass

(b)

Anisotropic rock mass

A careful selection of one of the above classes is essential for a reliable estimate
of the shear strength. Failure may take place along planar discontinuities when a
persistent discontinuity or a set of parallel discontinuities dip towards valley. An example
of a hillside failure along planar discontinuities is shown in Fig. 1. Kinematically, such a
failure will be possible if the joints strike parallel to the slope face and dip into the valley.
A more general and commonly occurring case is the one, when a pair of intersecting joint
planes forms a wedge that slides along the line of intersection of the joint planes.
The second class of rock shear strength pertains to the rock mass as a whole. This
class of shear strength is required in cases where the potential sliding surface lies partially
on discontinuity surfaces, and partially passes through the intact rock. The sliding surface
is generally curved just like in case of soils. Such failures occur when the extent of the
slope is very large compared to the discontinuity size and spacing. The rock mass
generally consists of blocks of intact rock separated by discontinuities. At the time of
slope failure, these blocks may slide, translate or rotate and dilation occurs as the normal
stresses are generally low. Further, the rock mass may behave isotropically or
anisotropically depending upon the number, orientation and spacing of discontinuities.
Though, there is no hard and fast rule, if a rock mass consists of more than four or five
sets of discontinuities, which are equally distributed in all directions, and no particular
discontinuity is substantially different from the others in shear strength response, then the

mass may be treated isotropic. On the other hand, if there are only about one to three sets
of discontinuities, the rock mass may be considered anisotropic.
The inherent anisotropy in the rock material itself is not a matter of serious
concern in rock slope problems, unless the failure is entirely through the intact rock. In
general, the stresses associated with rock slope stability problems are not that large to
cause failure of intact rock. Therefore, even if the intact rock material is inherently
anisotropic, the rock blocks forming the rock mass may be considered isotropic for all
practical purposes in rock slope problems.
1.2.

Some Important Terms


Having gone through the preliminary discussion in the previous section, it is

worth defining some of the terms, which will be used frequently in the subsequent text.
1.2.1.

Intact Rock
The rock material separated by discontinuities and free from any defect is termed

as the intact rock. Except in case of weak massive rock, intact rock failure generally does
not occur during failure of rock slopes. The strength of the intact rock is however,
required to assess the rock mass strength.
1.2.2.

Discontinuity
Any defect in the continuity of intact rock material is termed discontinuity. The

examples are bedding planes, joints and faults. The most common discontinuities
influencing rock slope stability are the joints. The term has sometimes been used
synonimically to cover all other types of the discontinuities. The discontinuities may be
tight or open, infilled or clear, smooth or rough. If the discontinuities are filled with a
material like clay or fault gouge, and there is no wall-to-wall contact, the shear strength
along the discontinuity will be governed by the characteristics of the infilling material.
The friction angle in such cases is likely to be low, and some cohesion may exist
depending on the type of infilling material. If the discontinuities are clean, the surfaces
will have low cohesion and the friction angle will depend on the roughness of the
material surface and bridges (gaps) between adjacent joints.

1.2.3.

Jointed Rock and Rock Mass


Rock mass is an assemblage of blocks of intact rock. Jointed rock is a general

term, which will be used in this text to indicate that the rock is fractured. It may not
necessarily represent a mass. For example, a laboratory cylindrical specimen having
single fracture will be termed as a jointed rock but not a rock mass. In case of failure of a
rock mass, intact rock blocks forming the mass may slide, translate and rotate. It results
in dilation of the mass if the normal stress level across the failure surface is low.
Consequently, the friction angle of mass at low normal stress is high, whereas the
cohesion is low. As the normal stress is increased, the dilation is suppressed and shearing
of the intact rock material commences. This results in a relatively higher cohesion and
lower friction angle. Due to continuous change in mechanism of failure, the strength
envelope of rock mass is highly curvilinear, especially in low normal stress range.
2. DISCONTINUITY SHEAR STRENGTH
If the probable mechanism of failure indicates a possibility of sliding along the
discontinuity surfaces (planar or wedge failure), it is essential to assess the shear strength
of discontinuity surfaces under the prevailing normal stress. Various models are available
to determine the shear strength of a planar discontinuity at a given normal stress. Some
most commonly referred shear strength models are discussed in the following text.
2.1.

Coulombs Model
The most commonly used model for assessing the shear strength along the

discontinuities is the conventional Coulombs linear model. In this model, shear strength
of the sliding surface is expressed in terms of cohesion, c j and friction angle, j. These
shear strength parameters may be obtained by performing direct shear tests on the
discontinuity surfaces. It is, however, difficult to prepare jointed rock test specimens of
regular geometry. A portable field shear box (Fig. 2a) as suggested by Ross-Brown and
Walton (1975), may be used to perform direct shear tests on rock joints. In this test, an
irregular rock chunk is encased in a cementing material in such a way that the
discontinuity surface coincides with the pre-defined shearing plane of the box. A normal
force is applied on to the discontinuity surface, and the specimen is sheared by subjecting

it to increasing shear force. Shearing force and shear displacements are monitored, and
the test carried to failure. A plot between shear stress and shear displacement is obtained.
A typical plot for a rough discontinuity tested under a constant normal load is presented
in Fig. 2b. The plot indicates that the shear displacement increases with increase in shear
stress. Initially the plot is linear indicating elastic behaviour of the specimen. Gradually
the force resisting movement is overcome; the curve becomes non-linear and reaches a
maximum that represents the peak shear strength of the discontinuity. With further
increase in shear displacement, the shear stress causing displacement decreases; and at
relatively large displacements, the shear stress reaches a constant value termed the
residual shear strength. The peak shear strength represents the maximum shear stress
required to overcome the resistance of a discontinuity before failure. After failure, the
cohesion of the discontinuity is lost; the minor irregularities on the rock surface are
sheared off, and further displacement requires relatively lesser shear stress.
A number of tests are conducted by adopting different values of normal loads. A
set of peak and residual shear strength values for different normal stresses will be
available from these tests. This data is used to obtain a plot between shear strength and
normal stress (Fig. 2c). The plot is termed as failure envelope. The diagram shows two
plots: one for peak strength and the other for residual strength. The shear strength
parameters cj, j for peak, and r for residual strength may be obtained from these plots.
The shear strength of the discontinuity is defined as:
f c j n tan

(peak strength)

(1)

f n tan r

(residual strength)

(2)

where f is the shear strength along the discontinuity; n is the effective normal
stress over the discontinuity; j is the peak friction angle of the discontinuity surface; cj is
the peak cohesion of the discontinuity surface, and r is residual friction angle for
discontinuity surface.
The above-mentioned expressions are for a rough or infilled discontinuity. If the
discontinuity is planar and clean, its cohesion will be zero and the shear strength will be
5

governed solely by the friction angle. The surface characteristics of rocks in such cases
will be governed by the size and shape of the grains exposed on the discontinuity surface
(Wyllie and Mah, 2004). A fine-grained rock with a high mica content aligned parallel to
the surface will tend to have low friction angle; while a coarse-grained rock such as
granite will have a high friction angle. A typical range of basic friction angles for a
variety of rocks is presented in Table 1.
2.2.

Pattons Model
Patton (1966) was probably one of the first research workers who recognized the

importance of surface roughness of natural discontinuities, quantified the roughness and


included its contribution to shear strength of natural discontinuities. Patton observed that
some bedded limestone slopes were stable at steeper angles whereas the other slopes
were unstable even at relatively gentle gradient. The reason for this anomaly was
attributed to the asperities (irregularities) on joint wall surfaces. The asperities were
divided into first order and second order asperities. At very low normal stress level, the
sliding is assumed to occur over second order asperities. With increase in normal stress,
the shearing of higher order asperities occurs, and the first order asperities define the
shearing strength. With further increase in normal stress, first order asperities are also
sheared. Laboratory studies were conducted (Patton, 1966) by simulating the asperities in
the form of saw-tooth specimens (Fig. 3). A bilinear model was suggested to predict the
shear strength as:
f = n tan ( +i)

for low n

(3)

f = c j n tan(r ) for high n

(4)

where i defines the roughness angle, and r is the residual friction angle.
Equation 3 is applicable for low normal stress levels where sliding occurs along
the asperities. At higher normal stress levels, shearing of intact rock occurs, and Equation
4 is suggested for obtaining the shear strength. The model, therefore, considers two

failure modes i.e. either sliding along the discontinuities or shearing of the intact
material.
The model provides a good basis for explaining the importance of roughness of
surfaces of the natural discontinuities and their influence on the slope stability. However,
from practical standpoint, the limitation of the model is that, it is difficult to assess the
normal stress level at which transition from sliding to shearing takes place. In reality,
there is no such distinct and clear-cut normal stress level, which defines the boundary
between the two failure modes. Due to this limitation, more advanced models have been
proposed which are discussed in the subsequent sections.
2.3.

Barton (1973) Model


Barton (1973) model is an extension of the Patton (1966) model. It is probably the

most widely used strength criterion for assessing shear strength along discontinuity
surfaces. It overcomes the limitations of Pattons model in that, all the parameters used in
the criterion can be easily determined in the field. Major limitation of Pattons model was
that it considered the shear failure either due to sliding or shearing. In nature, both the
mechanisms occur simultaneously. However, at low normal stress level, the sliding
phenomenon dominates; whereas shearing phenomenon governs the shear strength at
high normal stress level. The shear strength of a joint (Fig. 4) is expressed as:

JCS
f n tan r JRC log10

(5)

where JRC is the joint roughness coefficient, which is a measure of the initial
roughness (in degrees) of the discontinuity surface. JRC is assigned a value in the range
of 020, by matching the field joint surface profile with the standard surface profiles on a
laboratory scale of 10 cm (Barton and Choubey, 1977) as shown in Fig. 5. JCS is the joint
wall compressive strength of the discontinuity surface, and n is the effective normal
stress acting across the discontinuity surface.
Equation 5 may also be expressed in the following form:

f = n tan (e )

(6)

where e is the equivalent friction of the discontinuity at a given normal stress and
is expressed as:
e r i

<70

(7)
JCS

where roughness i = JRC log10

(8)

It can be seen that, the value of roughness angle, i goes on decreasing with
increasing n. At higher stress levels, when normal stress n equals JCS, the roughness
angle i, becomes zero. At this stress level, the equivalent friction angle, e of the
discontinuity, is equal to the basic friction angle, . At very low stress levels, the ratio
(JCS/n) tends to approach infinity. This may result in an unrealistic value of the
equivalent friction angle e. For practical purposes, this value should not exceed about
50o (Wyllie and Mah, 2004).
Joint roughness coefficient, JRC indicates initial roughness of the joint. This value
is obtained by comparing the profile of discontinuity surface with standard roughness
profiles, which have been prepared on a 10 cm base scale (Fig. 5). It has been found that,
with small displacements, the second order asperities are sheared off, and the first order
asperities come into picture. The roughness of joints is therefore a scale dependent
property. It is also found that JCS is also a scale dependent property. Barton and Bandis
(1982) proposed the following expressions to take into account the scale effect on the
roughness and wall strength of discontinuity surface:
L
JRCn = JRCo n
Lo
L
JCSn = JCSo n
Lo

0.02 JRCo

(9)

0.03 JRCo

(10)

where Lo is the dimension of the surface used to measure JRC and L n is the
dimension of sliding surface.
2.4.

Ladanyi and Archambault (1972) Criterion


Ladanyi and Archambault (1972) criterion is a mechanics based strength criterion

and has been derived from principles of strain energy. It is a criterion that is applicable to
both rock joints and jointed rock masses. To derive this criterion, the authors equated the
external work done in shearing a jointed rock to the internal energy stored in the rock. It
was considered that during shearing, the external force does the work to meet the energy
requirements for the following components:
i)

Component of work due to external work done in dilating against the external
force, N.

ii)

Additional internal work done in friction due to dilation only.

iii)

Work done in internal friction if the specimen did not change in volume in shear.

iv)

Work done in shearing asperities.


The criterion is expressed as:

n 1 a s v tan a s c i n tan i

(11)

1 1 a s v tan

where f is the shear strength; as is sheared area ratio equal to As/A; A is the total
area of joint surface; As is the sheared area of joint surface; n is the mean applied normal

is the rate of dilation at failure and is equal to


stress; v

dy
; is the basic friction angle
dx

joint surface, and ci, i are Mohr-Coulomb parameters for intact rock.

are difficult to estimate in the field.


The sheared area ratio, as and dilation rate, v
The authors have suggested on the basis of limited experimental data, the following
approximate expressions for determining these parameters:


a s 1 1 n
trn

(12)

tan i

(13)

K2

v 1 n
trn

K1

where K1 and K2 are equal to 1.5 and 4 respectively; trn is the brittle ductile
transition stress which may be taken equal to the UCS of intact rock; and i is the initial
roughness of the joints.
The criterion has been developed based on very sound principles of mechanics,
and is applicable to rock joints and rock masses as well, but lacks in simplicity due to the
fact that the parameters cannot be determined with ease in the field. Due to this reason,
the criterion could not become as popular amongst the practicing engineers, as the
empirical criterion suggested by Barton (1973).
3. SHEAR STRENGTH OF ROCK MASS
Geological conditions in many situations of rock slopes are such that the
discontinuity surfaces along which sliding could take place are not persistent. The
potential failure surface in such cases lies partly on discontinuity surface and partly
passes through the intact rock. The shear strength of rock mass is different from that of a
discontinuity and is very difficult to obtain experimentally, as the sample size would be
prohibitively large (> 1m diameter).
If the number of discontinuity sets in rock mass is large enough (say more than
four or so), and no particular discontinuity is significantly different from the others, and
spacing of discontinuities is small, the rock mass may be treated as isotropic. However, if
the number of joint sets is not large, the behaviour of rock mass will be anisotropic. If the
spacing of the discontinuities is large, it will be more appropriate to analyse the problem
by performing a wedge analysis. It may be realised that mobilised strength parameters of
rock mass are very low in rock slopes due to unrestricted rotation of blocks and freedom
to dilation in wedge failure in rock slopes, unlike that in tunnels (see article 3.3.2).

10

One of the most reliable methods to estimate rock mass shear strength is the back
analysis. It involves carrying out stability analysis of a failing or failed slope using
available information on different factors affecting the stability. A suitable value of factor
of safety is assigned and values of cohesion are calculated for a range of values of friction
angle. A realistic combination of shear-strength parameters c and is then selected for
analyzing slopes under similar geological conditions.
In the absence of back analysis, rock mass strength criteria are used to simulate
triaxial strength tests on the rock mass. Using the results of simulated triaxial strength
tests, a relationship between the normal stress across the failure surface and the
corresponding shear strength is derived. These relationships are generally non-linear.
Most of the non-linear strength criteria for rock masses are generally expressed in terms
of the major and minor principal stresses. Whereas, the stability analysis of the slopes is
generally carried out in terms of normal stress, shear stress and shear strength along the
failure surface. It is, therefore, necessary to convert strength values from (3, 1) space to
(,) space. Moreover, many standard procedures of slope stability problems e.g.
Bishops method (Singh and Goel, 2002) represent the solution in terms of MohrCoulomb shear strength parameters, c and . To use these procedures, it becomes
necessary to assess parameters c, at a given normal stress. It should be noted that the
failure envelope for rock masses is highly non-linear especially in low normal stress
range. To incorporate the non-linearity in shear strength behaviour, Hoek (2000) has
introduced the concept of instantaneous values of parameters, c and . The instantaneous
shear strength parameters, cisnt and inst are the values of cohesion intercept and angle of
the tangent drawn to the non-linear failure envelope at a point representing the normal
stress at which the shear strength is sought (Fig. 6). The instantaneous cohesion, c inst and
friction angle, inst are, therefore, normal stress dependent. Once the instantaneous
parameters are available, the shear strength may be obtained by using the linear
Coulombs model (Equation 1). The following expressions (Balmer, 1952) may be used
to convert the strength data from (3 ,1) space to (n ,f) space.
The strength criterion is expressed as:
11

1 = f(3)

(14)

then

n 3

1 3
1

1
3

3
f 1

1
1

tan

(15)

1
3

(16)

1
3

(17)

inst 2 90

(18)

cinst f n tan inst

(19)

where n is normal stress on failure plane; f is shear strength along the failure
plane; is the angle between the normal to failure plane and the major principal stress
direction; and inst, cinst are instantaneous Coulomb parameters. It should be noted that the
instantaneous shear strength parameters, cinst and inst vary with the level of normal stress
n acting on the failure plane.
3.1.

Linear Strength Criterion


Coulombs linear strength criterion is the most widely used criterion for jointed

rock and rock masses as well. The criterion is also referred to as Mohr-Coulomb
criterion. According to this criterion, rock mass is treated as an isotropic material and the
shear strength along the failure surface is expressed as follows:
f c j n tan j

(20)

12

where cj and j are Mohr-Coulomb shear strength parameters for jointed rock or
rock mass. For rock mass in field, the values of cj and j may be obtained from
classification approaches as given in the next section.
3.1.1.

Rock Mass Rating


Bieniawski (1973, 1989 and 1993) has suggested values of shear strength

parameters cm, m for five levels of rock mass ratings. Table 2 reproduces
recommendations of Bieniawski (1989) for cohesion and friction angle of rock mass
depending upon the rock mass rating, RMR. It can be seen that the values of c m have been
suggested to be varying from less than 100 kPa to more than 400 kPa and m values
varying from less than 15o to more than 45o respectively. Experience of Indian project
sites (Mehrotra, 1992) has however indicated that the shear strength is under predicted by
expressions suggested by Bieniawski (1989). Figure 7 (Mehrotra, 1992) shows results of
extensive block shear tests to get a relationship between RMR and shear strength
parameters. The figure shows that i) cm and m increase with increase in RMR, ii) values
obtained from field tests are higher than those suggested by Bieniawski (1989), iii) there
is a significant effect of the saturation of rock mass, and iv) friction angle varies nonlinearly with RMR. Further, for very poor rock masses (RMR = 0 20), friction angle, m
obtained from field test was always more than 15o whereas the same has been suggested
by Bieniawski (1989) to be less than 15o and approaching zero (or very small value) for
RMR 0. It has been suggested by Singh and Goel (1999) that for very poor rock
masses, the minimum value of m should always be taken greater than 15o. Figure 7 may
therefore be used for assessing the friction angle of rock masses.
A caution is needed here that there should be no double accounting for parameters
in the classification of rock mass and stability of a rock slope. Since ground water table
and joint orientations are considered in the stability analysis of slopes; the RMR should
be assessed on the basis of RQD, UCS, joint spacing, joint conditions and dry rock mass
(rating = +15). The strength parameters so obtained for the dry rock mass should be used
in stability analysis with water pressure being considered separately.

13

3.1.2.

Q index
Rock mass quality index, Q (Barton et al., 1974) can also be used to get the Mohr-

Coulomb shear strength parameters as suggested by Barton (2002) as:


RQD

Jn

c m

SRF

ci

100

MPa

Jr

J w 0.1
Ja

m tan 1

(21)

(22)

where cm is the cohesion of the undisturbed rock mass; m , the friction angle of
the mass; RQD, the rock quality designation (Deere, 1963); J n, the joint set number; Jr,
the joint roughness number; Ja, the joint alteration number; Jw, the joint water reduction
factor, and ci is the uniaxial compressive strength of intact rock material.
It may be noted that Q system has been used extensively and validated for tunnels
in Indian conditions. In the opinion of this author the shear strength parameters obtained
from Q should be preferred for analyzing underground openings. If used for slopes, an
overestimation in the strength may be expected. For slopes, it is felt that the relationship
between shear strength parameters and RMR as suggested by Mehrotra (1992), and
Singh and Goel (1999) will be more appropriate for Indian conditions and especially the
Himalayan rock masses. The relationships were developed based on extensive in-situ
direct shear tests on saturated rock masses in Himalayas.
3.2.

Non-Linear Strength Criteria


Mohr-Coulomb strength criterion considers the rock mass shear strength as a

linear function of normal stress n. It has been well established by now that at very low
normal stress level, which generally prevails in slopes, the shear strength response is
highly non-linear. Consequently, several non-linear strength criteria have been proposed
for jointed rock and rock masses to account for non-linearity in strength behaviour. In
general, these strength criteria have been presented in (1, 3) space, i.e. the major
principal stress at failure is expressed in terms of the minor principal stress. There is a

14

large number of rock strength criteria which are available in literature which include not
only non-linearity but also the effect of intermediate principal stress (Murrell,1963;
Weibols and Cook,1968; Mogi,1971; Chang and Haimson, 2000; Haimson and Chang,
2000; Colmenares and Zoback, 2002; and Al-Ajmi and Zimmerman, 2005). These criteria
are, however, beyond the scope of the present discussion.
From field point of view, only those criteria, whose parameters are easy to obtain
in the field, have been discussed in the subsequent sections. Due to simplicity and ease in
parameter estimation, the empirical failure criteria are widely used in practice. Only the
empirical criteria have, therefore, been presented herein. In the discussion that follows,
non-linear strength criteria have been grouped into two categories, i) strength criteria
applicable only to isotropic rock masses, and ii) the criteria applicable to anisotropic rock
masses as well. Under the first category Hoek-Brown criterion has been discussed and
under the second category Ramamurthy criterion and critical state concept based criterion
(Singh and Singh, 2004; Singh and Rao, 2005a) have been discussed.
3.2.1.

From Rock Mass Classification RMR or Q


Based on extensive in-situ direct shear testing of rock masses in the Himalayas,

Mehrotra (1992) has suggested the non-linear variation of shear strength as:

f
A n B
ci
ci

(23)

where A, B and C are empirical constants and depend on RMR or Q. Their values
for different moisture contents, RMR and Q index are presented in Table 3.
3.2.2.

Hoek-Brown Strength Criterion


A non-linear strength criterion was initially proposed for intact rocks by Hoek and

Brown (1980). The criterion is expressed as:


1 3 m i ci 3 ci2

(24)

15

where 1 is the effective major principal stress at failure; 3 is the effective


minor principal stress at failure; mi is a criterion parameter; and ci is the UCS of the
intact rock which is also treated as a criterion parameter.
A few triaxial tests may be conducted on the intact rock specimens, and the
criterion may be fitted into the triaxial test data to obtain ci and mi. It may be noted that
the computed value of ci should be used in further analysis and not the experimentally
obtained ci. In the absence of triaxial strength test data on intact rock, use of Table 4
(Hoek, 2000) has been suggested for approximate estimation of the parameter, mi.
The strength criterion proposed for intact rocks was also extended to rock masses
(Hoek and Brown, 1980). The criterion was later modified with more data available for
its applicability in the field for different rock types. The latest form of the criterion (Hoek
et al., 2002) is expressed as:

1 3 ci m j 3 s j
ci

(25)

where 1 is the effective major principal stress at failure; 3 is the effective


minor principal stress at failure; ci is the uniaxial compressive strength of intact rock
within the mass to be obtained from triaxial strength tests performed on intact rock
specimens; m j is an empirical constant which depends upon the rock type; and s j is an
empirical constant which varies between 0 to 1 depending upon the degree of fracturing.
To obtain parameter mj and sj, the use of a classification index, Geological
Strength Index (GSI), has been suggested (Hoek and Brown, 1997; Hoek et al., 2002).
The expressions for mj and sj are given as:
GSI 100
m j m i exp

28 14D

(26)

GSI 100
s j exp

9 3D

(27)

16

1 1 GSI / 15
e
e 20 / 3
2 6

(28)

where GSI is the Geological Strength Index; mi is the Hoek-Brown parameter for
intact rock to be obtained from triaxial test data; D is a factor which depends upon the
degree of disturbance to which the rock mass has been subjected by blast damage and
stress relaxation. It varies from 0 for undisturbed in situ rock masses to 1 for very
disturbed rock masses. For blasted rock slopes, D is taken in the range 0.7 to 1.0.
The Geological strength index (GSI) is assigned from visual inspection of the
rock mass in the field. The index depends upon the structure of the rock mass (massive,
blocky, very blocky, disturbed, seamy, disintegrated, laminated and sheared) and the
surface roughness conditions (very good, good, fair, poor and very poor) of
discontinuities. Like rock mass rating, RMR, the GSI values vary on a scale of zero to
one hundred. A chart for estimating GSI from Marinos et al., (2005) has been reproduced
in Fig. 8.
After assigning GSI, the Hoek-Brown parameters m j and sj are computed. These
parameters can then be used to simulate triaxial strength tests i.e., the 1 values can be
generated for various 3 as is done in the laboratory triaxial tests. From this data, the
failure envelope in vs. space can be obtained and Mohr-Coulomb parameters cinst and
inst may be determined at a given normal stress level. Alternatively Balmers expressions
(Balmer, 1952) may be used to generate f at a given n. Experience in Himalayan slopes
shows that cinst obtained from GSI, is too high in the case of rock slopes.
3.2.3.

Ramamurthy criterion
Ramamurthy

and

co-workers

(Ramamurthy,

1993;

Ramamurthy,

1994;

Ramamurthy and Arora, 1994; Ramamurthy, 2007) have suggested the following nonlinear strength criterion for intact isotropic rocks:
1 3
ci

B i
3 t
3 t

(29)

17

where 3 and 1 are the minor and major principal stresses at failure; t is the
tensile strength of intact rock; ci is the UCS of the intact rock; and i, i are the criterion
parameters.
Parameters i and Bi should be obtained by fitting the criterion into the laboratory
triaxial test data for intact rock. In the absence of triaxial test data, the following
approximate correlations may be used (Ramamurthy, 2007):

i 2 / 3; and B i 1.1 ci
t

1/ 3


to 1.3 ci
t

1/ 3

(30)

For jointed rocks and rock masses, the strength criterion proposed for intact rocks
has been extended to jointed rocks as:

1 3

B j cj
3
3

(31)

where j and j are the criterion parameters for jointed rock; and cj the is UCS of
the jointed rock.
Based on extensive laboratory testing, the following correlations were suggested
to obtain criterion parameters j and j:
j

cj

i ci

0.5

cj
Bi

0.13 exp 2.037


Bj

ci

(32)

0.5

(33)

where parameters i and i are obtained from laboratory triaxial tests performed
on intact rock specimens. The UCS of the rock mass, cj which is popularly known as
rock mass strength is an important input parameter to this strength criterion and has been
discussed separately in section 3.3.

18

3.2.4.

Critical state concept based criterion


As discussed in a preceding section, Mohr-Coulomb linear strength criterion is the

most widely used strength criterion in geotechnical engineering practice. Singh and Singh
(2004), and Singh and Rao (2005a) have suggested a strength criterion as an extension of
Mohr-Coulomb strength criterion. The linear Mohr-Coulomb strength criterion may be
expressed in the following form:

1 3 2c 0 cos 0
1 sin 0

1 sin 0
3
1 sin 0

(34)

where c0 and 0 are the instantaneous Mohr-Coulomb shear strength parameters at


30.
Figure 9 shows a plot of linear Mohr-Coulomb criterion. The actual response of a
rock mass is non-linear, which is also shown in the plot. The actual non-linear strength
response can be represented by modifying Mohr-Coulomb strength criterion as:

1 3 2c 0 cos 0
1 sin 0

1 sin 0
3 A ' 32 ; for 3 ci
1 sin 0

(35)

where A is an empirical constant, which is always positive.


Singh and Singh (2004), and Singh and Rao (2005a) have reduced the above
criterion into a simple form as:

1 3 cj 2Aci 3 A32 ; for 3 ci

(36)

where A is an empirical parameter, which is always, negative and is given as:


A = -1.23 (ci)-0.77

(37)

By using the values of rock mass strength, cj and intact rock strength, ci in the
criterion, a set of (3,1) values may be generated which then could be used to develop
failure envelope in (n,f) space. Alternatively, Balmers equations may be used in

19

stability analysis to determine instantaneous values of c and or shear strength for a


given normal stress level.
Rock Mass Strength, cj

3.3.

It can be observed from Equations 30, 31, 32 and 35, that rock mass strength, cj
is an important input parameter in the strength criteria. The estimation of rock mass
strength, therefore, becomes an important aspect in a rock slope problem. The accuracy
of shear strength prediction depends on how precisely cj has been estimated.
Ramamurthy criterion and parabolic criterion (Singh and Singh, 2004; Singh and Rao,
2005a) consider the rock mass to be anisotropic in strength response. The anisotropy is
introduced through rock mass strength, cj i.e. if an anisotropic value of rock mass
strength is used in the strength criterion, the resulting shear strength will also be
anisotropic. The following methods can be used to determine the UCS of the rock mass:
i)

Joint Factor concept

ii)

Rock mass quality, Q

iii)

Strength reduction factor

3.3.1.

Joint Factor Concept


Ramamurthy and co-workers (Arora, 1987; Ramamurthy, 1993; Ramamurthy and

Arora, 1994; Singh, 1997; Singh et al., 2002) have suggested that the presence of joints
greatly influences the rock mass strength. Joint Factor concept was evolved to define the
effect of joints on the strength of rock mass. The concept was derived from a large
number of uniaxial compressive strength tests conducted on jointed specimens of
artificial and natural rocks. Orientation of joints, their frequency and surface roughness
were varied during the tests. It was argued that the most important properties of joints
that affect the rock mass strength are frequency, orientation and shear strength along the
joints. An empirical parameter called joint inclination parameter, n (Table 5), was evolved
to represent the effect of orientation of joints on engineering response of jointed rocks. A
weakness coefficient, Joint Factor was defined by considering the combined effect of
frequency, orientation and shear strength of joints as:

20

Jf

Jn
n r

(38)

where Jn is the number of joints per metre in the major principal stress direction;
n is an inclination parameter that depends on the orientation of critical joints (Table 5); r
is the joint shear strength parameter, which depends on the conditions of the joint i.e.
roughness, cementation, tightness, aperture, weathering of walls, and nature of in-filled
material, if any (Ramamurthy, 2007).
The joint strength parameter, r is obtained from direct shear tests conducted along
the joint surface at low normal stress levels and is given by:
r

j
nj

tan j

(39)

where j is the shear strength along the joint; nj is the normal stress across the
joint surface; j is the equivalent value of friction angle incorporating the effect of the
asperities (Ramamurthy, 2007). The tests should be conducted at very low normal stress
levels so that the initial roughness is reflected through this parameter. For cemented
joints, the value of j includes the effect of cohesion intercept also. In case the direct
shear tests are not possible and the joint is tight, a rough estimate of j may be obtained
from Table 6 (Ramamurthy, 1994). If the joints are filled with gouge material and have
reached the residual shear strength, the value of r may be assigned from Table 7
(Ramamurthy, 1994).
The Joint Factor may be determined and the UCS of the rock mass (rock mass
strength) may be obtained as:
cj= ci exp(-0.008 Jf)

(40)

where cj is the uniaxial compressive strength of the jointed rock, and ci is the
UCS of intact rock. If there are more than one joint set available in the rock mass, J f

21

should be computed by considering the individual joint set and the maximum value
should be considered to assess the rock mass strength, cj.
The major part of the experimental work involved in deriving expression for rock
mass strength, cj in the above expression was performed on cylindrical jointed
specimens of size 38 cm diameter x 76 cm height. Singh (1997) performed an extensive
experimental study on relatively larger sized jointed blocky mass specimens of size 15
cm x 15 cm x 15 cm and having more that about 260 elemental blocks in each of the
specimens. It was observed that rock mass under uniaxial loading could fail due to the
following failure modes (Singh et al., 2002): i) splitting of intact material, ii) shearing of
intact material, iii) rotation of blocks and iv) sliding along the critical joints. An
assessment of the probable failure mode in the field can be made by using guidelines
suggested by Singh (1997) and Singh and Rao (2005a). The guidelines are reproduced in
the following section. It may be noted that in these guidelines, the joints are considered to
be continuous and persistent.
Let be the angle between the normal to the joint plane and the major principal
stress direction:
(i)

For = 0 to 10
The failure is likely to occur due to splitting of the intact material of blocks.

(ii)

For = 10 to 0.8 j
The mode of failure shifts from splitting (at = 10) to sliding (at 0.8 j).

(iii)

For = 0.8j to 65
The mode of failure is expected to be sliding only. Theoretically, the mass should

slide due to its own weight if > j; however, the experimental observations (Singh,
1997) indicate that the mass fails due to its own weight only if it has a single joint set. If
there are more joint sets, the mass deforms to some extent and becomes stable. This is

22

due to a small amount of rotation of blocks, which generates a small interlocking in the
mass. The net result is that the mass retains a small amount of strength.
(iv)

For = 65 to 75
The mode of failure shifts from sliding (at = 65) to rotation of blocks (at =

75).
(v)

For = 75 to 85
The mass fails due to rotation of blocks only. Geometry of the blocks is an

important parameter in governing the strength behaviour of the mass. In the present
study, it is assumed that the mass consists of blocks of square section. In case of slender
columns, the mass can fail due to buckling if the joints are open. Theory of long columns
can be used in this case and in the present analysis, this mode is excluded.
(vi)

For = 85 to 90
The failure mode shifts from rotation at = 85 to shearing at = 90. It may be

noted that sharp changes take place in the strength response in this range of orientations.
Table 8 has been suggested to roughly assess the failure mode in the field. The
table can be used for rock mass having one persistent joint set and a cross joint set.
Depending upon the orientation of persistent joints and the interlocking level of the mass,
a failure mode can be assigned from Table 8.
The following equation was suggested to determine the rock mass strength:
cj= ci exp(a Jf)

(41)

where a is an empirical multiplying constant depending upon the failure mode


(Table 9). If it is not possible to assess the failure mode, an average value of this
empirical constant may be taken -0.017.

23

Based on the assessed mode of failure, the value of coefficient a may be


assigned from Table 9. The value of Jf can be computed based on field observations (i.e.
frequency, orientation and friction angle along the joints). Equation 41 can now be used
to compute the rock mass strength.
The Joint Factor concept discussed above is quite useful in analysing anisotropic
strength behaviour of rock masses. For example, consider a case of a river valley with
joints dipping at some angle, and striking parallel to the river axis. For one bank, the
joints will dip towards valley; whereas on the other bank they will dip away from the
valley. Accordingly, the shear strength along the failure surface for a given normal stress
level, will be different for the two banks. That is why, on one side, a very steep slope may
be more stable due to large value of shear strength parameters whereas, the other bank
despite having gentle slope may be unstable due to low shear strength parameters.
3.3.2.

Rock Mass Quality, Q


Rock mass classification techniques have not only been used for classifying the

rock masses, but as design methodologies as well. Rock mass quality index, Q (Barton et
al., 1974) has been used extensively to classify rock masses. An estimate of rock mass
strength can also be made by using the rock mass quality index, Q. It should be noted that
the approach treats the rock mass as an isotropic medium.
Singh et al. (1997) have proposed correlations of rock mass strength, cj with Q
by analysing block shear tests in the field. The following expression has been suggested
to predict the rock mass strength for rock slopes in hilly areas:
cj 0.38Q1 / 3

MPa

(42)

where is the unit weight of rock mass in gm/cm 3; and Q is the Bartons rock
mass quality index.
It may be noted that this expression is applicable for rock slopes only. In case of
tunnels, the rock mass is subjected to a constrained dilatancy, which results in strength

24

enhancement and accordingly, the rock mass strength in tunnels is about 18 times larger
than that given by this expression (Singh and Goel, 2002).
3.3.3.

Strength Reduction Factor


Theoretically, the best estimates of rock mass strength, cj in the field can only be

made through large size field-testing in which the mass may be loaded upto failure to
determine rock mass strength. It is, however, extremely difficult, time consuming and
expensive to stress a large volume of jointed mass in the field upto ultimate failure. Singh
and Rao (2005b) have discussed that a better alternative is to get the deformability
properties of rock mass by stressing a limited area of the mass upto a certain stress level,
and then relate the ultimate strength of the mass to the laboratory UCS of the rock
material through a strength reduction factor, SRF. It has been shown by Singh and Rao
(2005b) that the modulus reduction factor, MRF and strength reduction factor, SRF are
correlated with each other by the following expression approximately:
SRF = (MRF)0.63
cj

E j

ci E i

(43)

0.63

(44)

where SRF is the strength reduction factor i.e. the ratio of rock mass strength to
the intact rock strength; MRF is the modulus reduction factor i.e. ratio of rock mass
modulus to the intact rock modulus; cj is the rock mass strength; ci is the intact rock
strength; Ej is the elastic modulus of the rock mass; and E i is the intact rock modulus
available from laboratory tests and taken equal to the tangent modulus at stress level
equal to 50% of the intact rock strength.
The elastic modulus of rock mass, Ej may be obtained in the field by conducting
uniaxial jacking tests (IS:7317, 1974). The test consists of stressing two parallel flat rock
faces on the opposite walls of a drift by means of a hydraulic jack (Mehrotra, 1992). The
stress is generally applied in two or more cycles as shown in Fig. 10. The second cycle of
the stress deformation curve is recommended for computing the field modulus as:

25

Ej

m(1 ) 2 P
A e

(45)

where Ej is the elastic modulus of the rock mass in kg/cm 2; is the Poissons
ratio of the rock mass (= 0.3); P is the load in kg; e is the elastic settlement in cm; A is
the area of plate in cm2; and m is an empirical constant =0.96 for circular plate of 25mm
thickness.
A number of methods for assessing the rock mass strength, cj have been
discussed above. It is desirable that more than one method be used for assessing the rock
mass strength and then generating the failure envelopes for the rock mass in - space. A
range of values will thus be obtained and design values may be taken according to
experience and confidence of the designer.
4. CONCLUSION
The stability of a rock slope is primarily governed by the characteristics of
discontinuities present in the mass. An appropriate understanding of shear strength along
the potential failure plane is essential to carry out limiting equilibrium analysis of rock
slopes. The potential failure plane may pass through existing joints or partly through the
joints and partly through the intact rock material. Based on potential failure mode, the
determination of shear strength has been divided into two broad categories i.e. joints and
jointed rocks / rock masses. Various approaches have been discussed to determine the
shear strength along the joints or jointed rock masses. Non-linear strength criteria, whose
parameters can easily be obtained in the field, have been discussed in detail. Both
isotropic and anisotropic strength responses of rock masses have been considered. An
important input parameter to the shear strength determination is the uniaxial compressive
strength of the rock mass or simply termed rock mass strength. Various recent methods to
determine the rock mass strength, developed from field and laboratory investigations,
have been discussed in detail. Using the strength criteria, one can generate shear strength
vs. normal stress plots for a potential failure surface in any direction, and use these results
in stability analysis. It is the opinion of the author that a range of values, rather than a
single value, of shear strength should be worked out to solve a real life problem in the
26

field. An appropriate value may then be chosen based on experience of the designer to
carry out the stability analysis. It may be realised that very low shear strength parameters
are mobilised in jointed rock mass slopes due to rotation of blocks.

ACKNOWLEDGEMENT
The author gratefully acknowledges the contributions made by Dr. T.
Ramamurthy, Professor (Retd.) and Dr. K.S. Rao Professor from IIT Delhi; Dr. Bhawani
Singh, Professor (Retd.), Prof. M. N. Viladkar and Prof. N. K. Samadhiya from IIT
Roorkee, Roorkee for their valuable technical inputs during the research presented in this
paper. Some part of the research presented in this paper has been conducted under a
research project (Project No. DST-209-CED, IIT Roorkee, 2005-2008) sponsored by
Department of Science and Technology (DST), New Delhi. The author sincerely puts on
record the appreciation for the financial support from DST New Delhi, and the cooperation and encouragement from Dr. Bhoop Singh, Director NRDMS, DST, New
Delhi, in carrying out research related to rock slope stability problems.
REFERENCES
Al-Ajmi Adel M. and Zimmerman R. W. (2005) Relation between the Mogi and the
Coulomb failure criteria, Int. Jl. Rock. Mech. Min. Sci., 42 , 431439.
Arora V.K. (1987) Strength and Deformational Behaviour of Jointed Rocks, Ph.D. Thesis,
IIT Delhi, India.
Balmer G. (1952) A general analytical solution for Mohr's envelope. Am. Soc. Test. Mat.,
52,1260-1271.
Barton N. (2002) Some new Qvalue correlations to assist in site characteristics and
tunnel design, Int. Jl. Rock Mech. Min. Sci., 39, 185 216.
Barton N.R. (1973) Review of a new shear strength criterion for rock joints. Engng Geol.
7, 287-332.
Barton N.R. and Bandis S.C. (1982) Effects of block size on the shear behaviour of
jointed rock. 23rd U.S. Symp. on Rock Mechanics, Berkeley, 739-760.

27

Barton N.R. and Choubey V. (1977) The shear strength of rock joints in theory and
practice. Rock Mech. 10 (1-2), 1-54.
Barton N.R., Lien R. and Lunde J. (1974) Engineering classification of rock masses for
the design of tunnel support. Rock Mech. 6(4), 189-239.
Bieniawski Z.T. (1973) Engineering classification of jointed rock masses. Transactions,
South African Institution of Civil Engineers, 15(12), 335-344.
Bieniawski Z.T. (1993) Classification of rock masses for engineering: The RMR system
and future trends. In: J.A. Hudson, Editor, Rock Testing and Site Characterization,
Pergamon Press, Oxford, UK, pp. 553573.
Bieniawski, Z.T. (1989) Engineering Rock Mass Classifications. New York: Wiley.
Chang C. and Haimson B. (2000) True triaxial strength and deformability of the German
continent deep drilling program (KTB) deep hole amphibolite, Jl. Gephys. Res.,
105(B8), 18999-013.
Colmenares L.B. and Zoback M.D. (2002) A statistical evaluation of intact rock failure
criteria constrained by polyaxial test data for five different rocks, Int. Jl. Rock Mech.
Min. Sci., 39, 695 -729.
Deere D.U. (1963) Technical description of rock cores for engineering purpose, Rock
Mechanics and Engineering Geology, 1, 18-22.
Haimson B. and Chang C. (2000) A new true triaxial cell for testing mechanical
properties of rock, and its use to determine rock strength and deformability of
Westerly granite, Int. Jl. Rock Mech. Min. Sci., 37, 285-296.
Hoek E. and Brown E.T. (1980) Empirical strength criterion for rock masses., Jl.
Geotech. Engng. Div, ASCE, 106(GT9), 1013-1035.
Hoek E. and Brown E.T. (1997) Practical estimates or rock mass strength. Int. Jl. Rock
Mech. & Min. Sci. 34(8), 1165-1186.
Hoek E. (2000) Practical Rock Engineering, 2000 Edition. http://www.rocscience.com/
roc/Hoek/Hoeknotes2000.htm.
Hoek E., Carranza-Torres C. and Corkum B. (2002) Hoek-Brown criterion 2002
edition. Proc. NARMS-TAC Conference, Toronto, 2002, 1, 267-273.
IS:7317 (1974) Code of Practice for uniaxial jacking test for modulus of deformation of
rocks.
Ladanyi B., Archambault G. (1972) Evaluation of shear strength of a jointed rock mass.
In: Proc. 24th Int. Geological Congress, Section 13D, pp. 249270. Montreal.

28

Marinos V., Marinos P. and Hoek E. (2005) The geological strength index: Applications
and limitations, Bull. Eng. Geol. Environ., 64: 5565.
Mehrotra, V.K. (1992) Estimation of Engineering Parameters of Rock Mass. PhD thesis.
University of Roorkee, Roorkee, India.
Mogi K. (1971) Fracture and flow of rocks under high triaxial compression. Jl. Geophys.
Res. 76:1255-69.
Murrell S.A.F. (1963) A criterion for brittle fracture of rocks and concrete under triaxial
stress, and the effect of pore pressure on the criterion. In: Fairhurst C, editor.
Proceedings of the 5th Symposium on Rock Mechanics, University of Minnesota,
Minneapolis, MN, 1963, 563-577.
Patton F.D. (1966) Multiple modes of shear failure in rock, Proc. Ist Cong. ISRM, Lisbon,
1, 509-513.
Ramamurthy T. (1993) Strength and modulus response of anisotropic rocks, In
Comprehensive Rock Engng., ed. J.A. Hudson, 313-329.
Ramamurthy T. (1994) Strength criterion for rocks with tensile strength, Proc. Indian
Geotechnical Conference, Warangal, 411-414.
Ramamurthy T. (2007) Strength, modulus and stress-strain responses of rocks,
Engineering in Rocks for Slopes, Foundations and Tunnels, Ed. T. Ramamurthy,
Prentice-Hall of India Pvt. Ltd., New Delhi, 93-137.
Ramamurthy T. and Arora V.K. (1994) Strength prediction for jointed rocks in confined
and unconfined states, Int. J. Rock Mech. Min. Sci., 13(1), 9-22.
Ross-Brown D. M. and Walton G. (1975) A Portable shear box for testing rock joints,
Rock Mechanics, 7, 129-153.
Singh B and Goel R.K. (2002) Software for Engineering Control of Landslide and Tunnel
Hazards, A.A. Balkema, The Netherlands, 344.
Singh B. and Goel R.K. (1999) Rock Mass Classification: A Practical Approach in Civil
Engineering, Elsevier Science Ltd., U.K. p.268.
Singh B., Viladkar M.N., Samadhiya N.K., Mehrotra V.K. (1997) Rock mass strength
parameters mobilised in tunnels. Tunn. Undergr. Space Technol. 12 (1), 4754.
Singh M. (1997) Engineering Behaviour of Jointed Model Materials, Ph.D. Thesis, IIT,
New Delhi, India.
Singh M. and Rao K.S. (2005a) Bearing capacity of shallow foundations in anisotropic
non Hoek-Brown rock masses, ASCE Journal of Geotechnical and Geoenvironmental Engineering, 131(8), 1014-1023.

29

Singh M. and Rao K.S. (2005b) Empirical methods to estimate the strength of jointed
rock masses, Engineering Geology, 77, 127-137.
Singh M. and Singh B. (2004) Critical state concept and a strength Criterion for rocks,
Proceedings 3rd Asian Rock Mechanics Symposium: Contribution of Rock Mechanics
to the New Century, eds. Ohinishi and Aoki, 877-880, Kyoto, Japan.
Singh M., Rao K.S. and Ramamurthy T. (2002) Strength and Deformational Behaviour of
Jointed Rock Mass, Rock Mech. & Rock Engng, 35(1), 45-64.
Wiebols G. A. and Cook N.G.W. (1968) An energy criterion for the strength of rock in
polyaxial compression. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr. 5, 529-49.
Wyllie D.C. and Mah C.W. (2004) Rock Slope Engineering, Spon Press, Taylor & Francis
Group, London and New York.

30

Вам также может понравиться