You are on page 1of 16

Chapter 26

Geological Strength Index

The function of Rock Mechanics engineers is not to compute accurately but to judge soundly.
Hoek and Londe

GEOLOGICAL STRENGTH INDEX


Hoek and Brown (1997) introduced the Geological Strength Index (GSI) for both hard
and weak rock masses. Experienced field engineers and geologists generally show a lik-
ing for a simple, fast, yet reliable classification that is based on visual inspection of geo-
logical conditions. A classification system should be non-linear for poor rocks as strength
deteriorates rapidly with weathering. Further, increased applications of computer model-
ing have created an urgent need for a classification system tuned to a computer simula-
tion of rock structures. To meet these needs, Hoek and Brown (1997) devised simple
charts for estimating GSI based on the following two correlations:
GSI ¼ RMR0 89  5 for GSI  18 or RMR  23 ð26:1Þ
0
¼ 9 lnQ þ 44 for GSI < 18 ð26:2Þ
0
where Q ¼ modified rock mass quality,
Q0 ¼ ½RQD=Jn½Jr=Ja, and ð26:3Þ
RMR0 89 ¼ rock mass rating according to Bieniawski (1989) when the groundwater
rating ¼ 15 and joint adjustment rating ¼ 0.
Sometimes, it is difficult to obtain RMR in poor rock masses, and Q0 may be used
more often because it is relatively more reliable than RMR, especially in openings in
weak rocks.
Hoek (Roclab, 2006) and Marinos and Hoek (2000) proposed a chart for GSI
(Figure 26.1) so experts can classify a rock mass by visual inspection alone. In this clas-
sification, there are six main qualitative rock classes, mainly adopted from Terzaghi’s
classification (Table 5.2).
1. Intact or massive
2. Blocky
3. Very blocky
4. Blocky/folded
5. Crushed
6. Laminated/sheared

Engineering Rock Mass Classification


# 2011 Elsevier Inc. All rights reserved. 319
320 Engineering Rock Mass Classification

FIGURE 26.1 Estimate of GSI based on visual inspection of geological conditions. (From Roclab,
2006; Marinos and Hoek, 2000) Modification by Cai et al. (2004) in terms of its quantification by block
volume and joint condition factor is also shown on the right side.

These classifications have been available to engineers and geologists for 60 years. Dis-
continuities are classified into five surface conditions that are similar to joint conditions
in RMR (Chapter 6).
1. Very good
2. Good
Chapter 26 Geological Strength Index 321

3. Fair
4. Poor
5. Very poor
A 6  5 block in the matrix of Figure 26.1 is picked up first according to actual and
undisturbed rock mass classification and discontinuity surface condition. Then a corre-
sponding GSI is read. According to Hoek (1998) and Marinos and Hoek (2000), a range
of values of GSI (or RMR) should be estimated instead of just a single value. This prac-
tice has a significant impact on the design of slopes and excavations in rocks. Drastic
degradation in GSI, RMR, and Q-values is found to occur in openings after squeezing
and rock bursts. This is also seen in openings, hence the need for evaluating the GSI
of rock mass in the undisturbed condition (D ¼ 0). Back analysis of both a model
(polyaxial strength criterion) and its parameters (from the observed behavior of rock
structures) is an ideal method of the rock mass characterization, and GSI is the first step
in this direction.
Figure 26.1 is used judiciously for crushed/disintegrated and laminated/sheared
rocks. Similarly, hard, thick laminated rocks in the last row of Figure 26.1 may
not be applicable, because they may have a higher strength classification (see
Table 5.2, Class II).
The GSI chart has been subsequently quantified by Cai et al. (2004) by incorpo-
rating the rock block volume (Vb) formed by the joints or discontinuities and the joint
condition factor JC (see Table 4.6). The suggested quantification is also shown in
Figure 26.1. The block volume (Vb), affected by the joint set spacing and
persistence, can broadly be known by the joint spacing given for six different rock
classes in Figure 26.1. The value of joint condition factor, JC, controlled by joint
roughness, weathering, and infilling material, can be obtained by Eq. (26.4) from
Cai et al. (2004).
JW JS
JC ¼ ð26:4Þ
JA
where JW ¼ large-scale joint or discontinuity waviness in meters from 1 to 10 m
(Table 26.1), JS ¼ small-scale smoothness in centimeters from 1 to 20 cm (Table 26.2),
and JA ¼ joint alteration factor (Table 26.3).
Cai and Kaiser (2006), based on the proposed quantitative chart (Figure 26.1), and
using surface fitting techniques, suggested the following equation to calculate GSI from
JC and Vb:
26:5 þ 8:79 ln JC þ 0:9 lnVb
GSIðVb , JC Þ ¼ ð26:5Þ
1 þ 0:0151 ln JC  0:0253 lnVb
where JC is a dimensionless factor defined by Eq. (26.4) and block volume Vb is in cm3
(see the section Calibration of RMi from Known Rock Mass Strength Data in
Chapter 10).
To avoid double-accounting, groundwater condition and in situ stresses are not con-
sidered in GSI because they are accounted for in computer models. GSI assumes that
the rock mass is isotropic; therefore, only cores without weak planes should be tested
in triaxial cells to determine qc and mr as GSI downgrades strength according to schis-
tocity. This classification reduces many uncertainties in rock mass characterization. An
undisturbed rock mass should be inspected for classification; however, heavy blasting
creates new fractures.
322 Engineering Rock Mass Classification

TABLE 26.1 Terms to Describe Large-Scale Waviness (JW)

Rating for
Waviness waviness
terms Undulation (JW)
Interlocking 3
(large-scale)

Stepped 2.5
Large >3% 2
undulation
Small to 0.3–3% 1.5
moderate
undulation
Planar <0.3% 1

Sources: Palmstrom, 1995; Cai et al., 2004.

TABLE 26.2 Terms to Describe Small-Scale Smoothness (JS)

Smoothness Rating for


terms Description smoothness (JS)
Very rough Near vertical steps and ridges occur with 3
interlocking effect on the joint surface
Rough Some ridges and side-angles are evident; asperities 2
are clearly visible; discontinuity surface feels
very abrasive (rougher than sandpaper grade 30)
Slightly rough Asperities on the discontinuity surfaces 1.5
are distinguishable and can be felt
(like sandpaper grade 30–300)
Smooth Surface appears smooth and feels so to the touch 1
(smoother than sandpaper grade 300)
Polished Visual evidence of polishing exists; this is often 0.75
seen in coating of chlorite and especially talc
Slickensided Polished and striated surface that results from 0.6–1.5
sliding along a fault surface or other movement
surface

Sources: Palmstrom, 1995; Cai et al., 2004.


Chapter 26 Geological Strength Index 323

TABLE 26.3 Rating for Joint Alteration Factor (JA)

Term Description JA

Rock wall Clear joints


contact
Healed or “welded” joints Softening, impermeable filling 0.75
(unweathered) (quartz, epidote, etc.)
Fresh rock walls No coating or filling on joint 1
(unweathered) surface, except for staining
Alteration of joint wall: The joint surface exhibits 2
slightly to moderately one class higher alteration
weathered than the rock

Alteration of joint wall: The joint surface exhibits 4


highly weathered two classes higher alteration
than the rock
Coating or thin filling
Sand, silt, calcite, talc, etc. Coating of frictional material 3
without clay
Clay, chlorite, talc, etc. Coating of softening and 4
cohesive minerals
Filled joints Sand, silt, calcite, etc. Filling of frictional material 4
with partial without clay
or no
contact Compacted clay materials “Hard” filling of softening 6
between and cohesive materials
the rock Soft clay materials Medium to low over- 8
wall consolidation of filling
surfaces
Swelling clay materials Filling material exhibits 8–12
swelling properties

Sources: Palmstrom, 1995; Cai et al., 2004.

GENERALIZED STRENGTH CRITERION


Hoek, Carranza-Torres, and Corkum (2002) suggested the following generalized
Hoek-Brown strength criterion for undisturbed rock masses:
s3
s1 ¼ s3 þ qc ½mb þ sn ð26:6Þ
qc
where s1 ¼ maximum effective principal stress, s3 ¼ minimum effective principal
stress, qc ¼ uniaxial compressive strength (UCS) of rock material (intact) for standard
NX size core (see Table 8.13 after Palmstrom, 2000), mb ¼ reduced value of the material
constant mr, and
 
GSI  100
mb ¼ mr  exp ð26:7Þ
28  14D
324 Engineering Rock Mass Classification

where mr ¼ Hoek-Brown rock material constant to be found from triaxial tests on


rock cores.
In Eqs. (26.8) and (26.9), s and n are Hoek-Brown constants for the rock mass given
by the following relationships:
 
GSI  100
s ¼ exp ð26:8Þ
9  3D
1 1 
n ¼ þ eGSI=15  e20=3 ð26:9Þ
2 6
D is a disturbance factor that 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 (Table 26.4).
Cheng and Liu (1990) found that a zone of blast damage extended for a distance of
approximately 2.0 m with D ¼ 0.7 around all large excavations (caverns). While using
the disturbance factor D, its values given in Table 26.4 are selected judiciously. The
actual value of D is a function of rock mass quality and blasting practices.
Experience in the design of slopes in very large open pit mines has shown that the
Hoek-Brown criterion for undisturbed in situ rock masses (D ¼ 0) results in shear
strength parameters that are too optimistic. The effects of heavy blast damage as well

TABLE 26.4 Guidelines for Estimating Disturbance Factor D

Suggested
Appearance of rock mass Description of rock mass value of D
Excellent quality controlled blasting or D¼0
excavation by tunnel boring machine
results in minimal disturbance to the
confined rock mass surrounding a tunnel.

Mechanical or hand excavation in poor D¼0


quality rock masses (no blasting) results in D ¼ 0.5
minimal disturbance to the surrounding No invert
rock mass.
Where squeezing problems result in
significant floor heave, disturbance can be
severe unless a temporary invert, as shown
in the photograph, is placed.
Chapter 26 Geological Strength Index 325

TABLE 26.4—Cont’d

Suggested
Appearance of rock mass Description of rock mass value of D
Very poor quality blasting in a hard rock D ¼ 0.8
tunnel results in severe local damage,
extending 2 or 3 m, in the surrounding
rock mass.

Small-scale blasting in civil engineering D ¼ 0.7


slopes results in modest rock mass damage, Good
particularly if controlled blasting is used blasting
as shown on the left-hand side of the D ¼ 1.0
photograph. However, stress relief Poor blasting
results in some disturbance.

Very large open pit mine slopes suffer D ¼ 1.0


significant disturbance due to heavy Production
production blasting and also due to stress blasting
relief from overburden removal. D ¼ 0.7
In some softer rocks, excavation can be Mechanical
carried out by ripping and dozing and excavation
the degree of damage to the slopes is less.

Sources: Hoek et al., 2002; Hoek, 2007.

as stress relief due to removal of the overburden of the rock mass results in disturbance of
the rock mass. It is considered that the “disturbed” rock mass parameters with D ¼ 1 in
Eqs. (26.7) and (26.8) are more appropriate for these rock masses (Hoek et al., 2002).
Thus, UCS of a rock mass obtained from Eq. (26.6) is
qcmass ¼ qc  sn ð26:10Þ
and uniaxial tensile strength (UTS) of a good rock mass is
sq
qtmass ¼  c ð26:11Þ
mb
326 Engineering Rock Mass Classification

Equation (26.11) is obtained by setting s1 ¼ s3 ¼ qtmass in Eq. (26.6). This represents a


condition of biaxial tension. Hoek (1983) showed that the UTS is equal to the biaxial
tensile strength for brittle materials.
Hoek (2007) proposed Eq. (13.5) for estimating rock mass strength (qcmass) from lab-
oratory strength of intact rock material (qc) and GSI for D ¼ 0.

MOHR-COULOMB STRENGTH PARAMETERS


Mohr-Coulomb’s strength criterion for a rock mass is expressed as
s1  s3 ¼ qcmass þ As3 ð26:12Þ
where qcmass ¼ UCS of the rock mass, which ¼ 2 c cosf/(1  sinf); c ¼ cohesion of the
rock mass; A ¼ 2 sinf/(1  sinf); and f ¼ angle of internal friction of the rock mass.
Hoek and Brown (1997) made extensive calculations on the linear approximation of
non-linear strength criterion (Eq. 26.6). They found that strength parameters c and f de-
pend upon s3; thus, they plotted charts for average values of c (Figure 26.2) and f
(Figure 26.3) with D ¼ 0 for a quick assessment. It may be noted that c and f decrease
non-linearly with GSI unlike RMR (Table 6.10). The rock parameter mr may be guessed
from fp of a rock material at GSI of 90, if adequate triaxial tests are not done. Table 26.5
lists typical values of mr for various types of rock materials.
The angle of dilatancy of a rock mass after failure is recommended approximately as
D ¼ ðf=4Þ for GSI ¼ 75
¼ ðf=8Þ for GSI ¼ 50 ð26:13Þ
¼0 for GSI  30

FIGURE 26.2 Relationship between ratio of cohesive strength of rock mass to UCS of intact rock (c/qc)
and GSI for different mr values for D ¼ 0. (From Hoek and Brown, 1997)
Chapter 26 Geological Strength Index 327

FIGURE 26.3 Friction angle (f) of rock mass for D ¼ 0 for different GSI and mr values. (From Hoek
and Brown, 1997)

The Hoek et al. (2002) correlations for s are valid for rock slopes and open pit mines, but
not for structurally controlled rock slopes and transported rockfill slopes. For tunnels and
caverns, there is an enormous strength enhancement (Chapter 13).

MODULUS OF DEFORMATION
Hoek and Diederichs (2006) found a useful correlation for modulus of deformation (Ed)
of rock mass based on approximately 496 in situ tests.
 
1  D=2
Ed ¼ Er 0:02 þ , GPa ð26:14Þ
1 þ expðð60 þ 15D  GSIÞ=11Þ
where Er ¼ modulus of elasticity of intact rock in GPa.
The elastic modulus (Ee) is obtained from the unloading cycles of the uniaxial jacking
tests. It is correlated for both dry and saturated rock mass as follows (Chapter 8 and
Eq. 8.19):
Ee ¼ 1:5Q0:6 E0:14
r , GPa ð26:15Þ
where Q ¼ rock mass quality.
Equation (26.15) is suggested for the dynamic analyses of concrete dams during a
major earthquake and machine (generator) foundations on the rock masses.
The original equation proposed by Hoek and Brown (1997) has been modified by the
inclusion of factor D to allow for the effects of blast damage and stress relaxation. The
strength and deformation parameters estimated from the GSI system are very close to
those obtained from in situ tests (Cai et al., 2004). Back analysis of observed displace-
ments in openings may give more realistic values of the design parameters including the
disturbance factor by trial and error.
328 Engineering Rock Mass Classification

TABLE 26.5 Values of the Constant mr for Intact Rock Material by Rock Group

Texture

Rock type Class Group Coarse Medium Fine Very fine


Sedimentary Clastic Conglomerate Sandstone Siltstone Claystone
(22) 19 9 4

Greywacke
(18)

Non- Organic ---------- Chalk ----------


clastic 7
---------- Coal ----------
(8–21)
Carbonate Breccia Sparitic Micritic —
(20) limestone (10) limestone
8
Chemical — Gypstone Anhydrite —
16 13
Metamorphic Non-foliated Marble Hornfels Quartzite —
9 (19) 24
Slightly foliated Migmatite Amphibolite Mylonites —
(30) 25–31 (6)

Foliated* Gneiss Schists Phyllites Slates


33 4–8 (10) 9

Igneous Light Granite — Rhyolite Obsidian


33 (16) (19)
Granodiorite — Dacite —
(30) (17)
Diorite — Andesite —
(28) 19
Dark Gabbro Dolerite Basalt —
27 (19) (17)
Norite — — —
22
Extrusive Agglomerate Breccia Tuff —
pyroclastic type (20) (18) (15)

The values given are estimates. It is suggested to get the mr values from triaxial test data.
*These values are for intact rock specimens tested normal to bedding or foliation. The value of mr will be
significantly different if failure occurs along a weakness plane.
Source: Hoek, Marinos, and Benissi, 1998.
Chapter 26 Geological Strength Index 329

ROCK PARAMETERS FOR INTACT SCHISTOSE


In argillaceous or anisotropic rocks (shales, phyllites, schists, gneisses, etc.), the UCS of
rock material qc depends upon the orientation of the plane of weakness. Both GSI and
RMR take into account the orientation of joints. To avoid double-accounting for joint
orientation in both UCS and GSI, it is a common engineering practice to use the upper
bound value of qc and corresponding mr for rock cores with nearly horizontal planes of
weakness for estimating mb, s, and Ed for jointed rock masses.
Cohesion along joints is needed for wedge analysis or computer modeling. Cohesion
along bedding planes or planar continuous joints (longer than 10 m) may be negligible.
However, cohesion along discontinuous joints (assumed continuous in the wedge anal-
ysis) may be the same as cohesion (c) of the rock mass. The cohesion of the rock mass is
due to the cohesion of the discontinuous joints. The ratio of c and cohesion of rock ma-
terial (Figure 26.2) may be of the same order as the area of intact rock bridges per unit
area of discontinuous joints.

ESTIMATION OF RESIDUAL STRENGTH OF ROCK MASSES


To extend the GSI system for estimation of rock mass residual strength, Cai et al. (2007)
proposed an adjustment of the original GSI value based on the two major controlling fac-
tors in the GSI system, block volume (Vb) and joint condition factor (JC), to reach the
residual values.
The difference between the peak and residual strength of a rock mass with non-
persistent joints is larger than that of a rock mass with persistent joints. The implication
is that a drop of GSI from peak to residual values is larger for rock masses with non-
persistent joints. Besides rock bridges, rock asperity interlocking also contributes to
the difference between peak and residual strengths.

Residual Block Volume


If a rock experiences post-peak deformation, the rock in the broken zone is fractured
and consequently turned into a poor and eventually “very poor” rock (Figure 7.2).
The properties of a rock mass after extensive straining should be derived from the rock
class of “very poor rock mass” in the RMR system (Chapter 6) or “disintegrated” in the
GSI system.
For the residual block volume, it is observed that the post-peak block volume is small
because the rock mass has experienced tensile and shear fracturing. After the peak load,
the rock mass becomes less interlocked and is heavily broken with a mixture of angular
and partly rounded rock pieces.
Detailed examination on the rock mass damage state (before and after the in situ
block shear tests at some underground cavern sites in Japan) revealed that in areas
not covered by concrete, the failed rock mass blocks are 1–5 cm in size. The rock mass
is disintegrated along a shear zone in these tests. As such, Cai et al. (2007) suggested the
following residual block volume Vbr :
l If Vb > 10 cm3, Vbr (in disintegrated category) ¼ 10 cm3
l If Vb < 10 cm3, Vbr ¼ Vb
330 Engineering Rock Mass Classification

Residual Joint Condition Factor


The residual joint surface condition factor JCr is calculated now from Eq. (26.16).
JrW JrS
JrC ¼ ð26:16Þ
JAr
r
where JW , JSr, and JA
r
are residual values of large-scale waviness, small-scale smooth-
r
ness, and joint alteration factor, respectively. The reduction of JW and JSr is based on
the concept of mobilized joint roughness and the equations are given as
JW JW
If < 1, JrW ¼ 1; Else JrW ¼ ð26:17Þ
2 2
JS JS
If < 0:75, JrS ¼ 0:75; Else JrS ¼ ð26:18Þ
2 2
There is no reduction in JA.

Residual GSI Value and Strength Parameters


The residual GSIr is a function of Vbr and JCr, which can be estimated using Eq. (26.5).
Fracturing and shearing do not weaken the intact rocks (even if they are broken into
smaller pieces) so the mechanical parameters (qc and mr) should be unchanged. There-
fore the generalized non-linear criterion for the residual strength of jointed rock masses
can be written as
 nr
s3
s1 ¼ s3 þ qc mbr þ sr ð26:19Þ
qc
where mbr, sr, and nr are the residual constants for the rock mass. These constants can be
determined from a residual GSIr (Cai et al., 2007).
 
GSIr  100
mbr ¼ mr  exp ð26:20Þ
28
 
GSIr  100
sr ¼ exp ð26:21Þ
9
1 1 
nr ¼ þ eGSIr =15  e20=3 ð26:22Þ
2 6
Because the rock masses are in a damaged, residual state, D ¼ 0 is used for the residual
strength parameter calculation.

CLASSIFICATION OF SQUEEZING GROUND CONDITION


Hoek (2001) classified squeezing ground conditions based on tunnel strain (ua/a) or
the ratio between rock mass strength and in situ stress (gH), as shown in Figure 26.4.
In very severe squeezing ground (ua/a >5%), the tunnel face may exhibit plastic extru-
sion due to the failure of rock mass all around the tunnel and the face has to be stabilized.
For a rock mass strength (qcmass) of 1.5 MPa and in situ stress of 13.5 MPa (gH), the ratio
(qcmass/gH) ¼ 0.11. Figure 26.4 shows that this corresponds to a tunnel strain of approx-
imately 10% and very severe squeezing ground condition should be anticipated.
Chapter 26 Geological Strength Index 331

FIGURE 26.4 Tunneling problems associated with different levels of strain. (From Hoek, 2001; Singh
and Goel, 2006)

Example 26.1
In a major hydroelectric project in dry quartzitic phyllite, the rock mass quality (Q) is in
the range of 6 to 10. The joint roughness number Jr is 1.5 and joint alteration number Ja is
1.0 for critically oriented joints in the underground machine hall. The unit weight of
phyllite rock is 2.78 gm/cc. The upper bound strength envelope between s1 and s3 from
triaxial tests gave UCS (qc) ¼ 80 MPa, fp ¼ 32 , mr ¼ 5.3, and Er ¼ 11.6 GPa when the
plane of schistocity is horizontal. The average UCS for various angles of schistocity is
40 MPa. The GSI is estimated to be about 55 as rock mass is micro-folded and joints
are very rough and unweathered. With these values, it is required to consider the engi-
neering parameters of the undisturbed (D ¼ 0) rock mass for the machine hall cavity
(width 24 m and height 47 m).
The average rock mass quality is √(6  10) ¼ 8 (approximately). Other calculations
are presented in Table 26.6 for the undisturbed rock mass. The peak angle of internal
friction works out to be 27 from Figure 26.3 and 32 from triaxial tests and 56 from
the Jr/Ja value. Thus, a value of fp ¼ 32 appears to be realistic. A blast damaged zone
of about 2 m depth may be assumed in the computer modeling all around the cavity with
half the values of cp, qcmass, Ed, and G.
It may be emphasized that Table 26.6 suggests parameters for the first iteration only
in the computer modeling. The more realistic model and parameters may be back cal-
culated from the observed displacements of the cavity during upper half excavation.
332

TABLE 26.6 Recommended Engineering Parameters of Undisturbed Rock Mass

S. No. Rock mass parameter Reference Recommended value Remarks


1 n Eq. (26.9) 0.5 —
mb Eq. (26.7) 1.1 D¼0
3
s Eq. (26.8) 6.7  10 D¼0
2 cp Figure 26.2 3.6 MPa qc ¼ 80 MPa

3 fp — 32 Same as that of rock material
4 UCS qcmass 2 cp cosfp /(1  sinfp) 13 MPa Intercept on s1 and s3 envelope
5 UTS qtmass 0.029 g Q 0.31
0.15 MPa Q¼8

6 Angle of dilatancy D (fp  fr)/2 5 —
 
7 fr fp  10  14 22 —
8 Residual cohesion cr Chapter 13; see the section Tensile 0.1 MPa —
Strength Across Discontinuous Joints
9 Residual UCS 2 cr cosfr /(1  sinfr) 0.3 MPa —
10 Modulus of deformation Ed Uniaxial jacking test 7.5 MPa Pressure dependency not observed
11 Poisson’s ratio — 0.20
12 Shear modulus Ed/10 0.75 MPa Axis of anisotropy along bedding
plane
13 Suggested model for peak Eq. (13.14) 13 þ 2.2(s2 þ s3)/2 MPa —
strength
14 Model for residual strength Mohr-Coulomb’s theory 0.3 þ 1.2 s3 MPa
Chapter 26 Geological Strength Index 333

Example 26.2
Given the strength of rock material (qc) ¼ 50 MPa, Hoek-Brown parameters for rock
material (mr) ¼ 10, GSI ¼ 45, and overburden above tunnel (H) ¼ 100 m. Estimate
the shear strength parameters of both undisturbed and heavily blasted rock mass
(D ¼ 1.0) using Hoek’s computer program Roclab (2006).
For an undisturbed in situ rock mass surrounding a tunnel at a depth of 100 m, with a
disturbance factor D ¼ 0, the equivalent friction angle f ¼ 47 while the cohesive
strength is c ¼ 0.58 MPa. A rock mass with the same basic parameters but in a highly
disturbed slope of 100 m height, with a disturbance factor of D ¼ 1, has an equivalent
friction angle of f ¼ 28 and a cohesive strength of c ¼ 0.35 MPa.

REFERENCES
Bieniawski, Z. T. (1989). Engineering rock mass classifications (p. 251). New York: John Wiley.
Cai, M., & Kaiser, P. K. (2006). Visualization of rock mass classification systems. Geotechnical and Geo-
logical Engineering, 24(4), 1089–1102.
Cai, M., Kaiser, P. K., Tasaka, Y., & Minami, M. (2007). Determination of residual strength parameters of
jointed rock masses using GSI system. International Journal of Rock Mechanics and Mining Sci-
ences, 44, 247–265.
Cai, M., Kaiser, P. K., Uno, H., Tasaka, Y., & Minami, M. (2004). Estimation of rock mass deformation
modulus and strength of jointed hard rock masses using the GSI system. International Journal of Rock
Mechanics and Mining Sciences, 41, 3–19.
Cheng, Y., & Liu, S. (1990). Power caverns of the Mingtan pumped storage project, Taiwan. In J. A.
Hudson (Ed.), Comprehensive Rock Engineering (Vol. 5, pp. 111–132). Oxford, UK: Pergamon.
Hoek, E. (1983). Strength of jointed rock masses. 23rd Rankine Lecture. Institution of Civil Engineers.
Geotechnique, 33(3), 187–223.
Hoek, E. (1998). Reliability of Hoek-Brown estimates of rock mass properties and their impact on design.
Technical Note. International Journal of Rock Mechanics and Mining Sciences, 35(1), 63–68.
Hoek, E. (2001). Big tunnels in bad rock, 36th Terzaghi Lecture. Journal of Geotechnical and Geo-
environmental Engineering, ASCE, 127(9), 725–740. http://150.217.9.3/geotecnica/hoek_badrock.pdf.
Hoek, E. (2007). Practical rock engineering. Chap. 12. www.rocscience.com.
Hoek, E., & Brown, E. T. (1997). Practical estimates of rock mass strength. International Journal of Rock
Mechanics and Mining Sciences, 34(8), 1165–1186.
Hoek, E., Carranza-Torres, C., & Corkum, B. (2002). Hoek-Brown Failure Criterion — 2002 edition.
In 5th North American rock mechanics Symposium (Vol. 1, pp. 267–273). 17th Tunnel Association
of Canada, NARMS-TAC Conference, Toronto.
Hoek, E., & Diederichs, M. S. (2006). Empirical estimation of rock mass modulus. International Journal
of Rock Mechanics and Mining Sciences, 43, 203–215.
Hoek, E., Marinos, P., & Benissi, M. (1998). Applicability of the Geological Strength Index (GSI) clas-
sification for very weak and sheared rock masses—The case of Athens schist formation. Bulletin of
Engineering Geology and Environment, 57, 151–160.
Marinos, P., & Hoek, E. (2000). GSI — A geologically friendly tool for rock mass strength estimation.
In Proceedings of the GeoEngineering 2000 Conference. Melbourne, Australia.
Palmstrom, A. (1995). RMi — A system for characterising rock mass strength for use in rock engineering.
Journal of Rock Mechanics and Tunnelling Technology, 1(2), 69–108.
334 Engineering Rock Mass Classification

Palmstrom, A. (2000). Recent developments in rock support estimates by the RMi. Journal of Rock
Mechanics and Tunnelling Technology, 2(1), 1–24.
Roclab, A. (2006). Computer program ‘Roclab’ downloaded from Rocscience web site. www.rocscience.
com.
Singh, B., & Goel, R. K. (2006). Tunnelling in weak rocks (p. 488). Amsterdam: Elsevier.

You might also like