Bien Iaw Ski 1993

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

22

Classification of Rock Masses for


Engineering: The RMR System and
Future Trends
z.T. BIENIAWSKI
Pennsylvania State University, University Park, PA, USA

22.1 INTRODUCTION 553


22.2 PURPOSES OF ROCK MASS CLASSIFICATION 554
22.3 CLASSIFICATION PARAMETERS 554
22.4 MAJOR CLASSIFICATIONS CURRENTLY IN USE 555
22.5 THE ROCK MASS RATING SYSTEM 555
22.5.1 Definition of the RMR System 556
22.5.2 Classification Procedure 556
22.5.3 Applications of the RMR System 565
22.5.3.1 Rock mass deformability 565
22.5.3.2 Rock mass strength 566
22.5.3.3 Advantages and limitations 568
22.5.3.4 An example of use 569
22.5.4 Database/or the RMR System 569
22.5.5 Correlations with Other Classification Systems 570
22.6 FUTURE TRENDS 572
22.7 SUMMARY AND CONCLUSIONS 572
22.8 REFERENCES 572

22.1 INTRODUCTION
Rock mass classifications form the backbone of the empirical design approach which relates
practical experience gained on previous projects to the conditions anticipated at a proposed site.
They are widely employed in rock engineering [1]. In fact, on many projects, the classification
approach serves as the only systematic and practical basis for the design of complex excavations in
rock. Most of the tunnels, mines, slopes and foundations constructed currently make use of
a classification system.
The best known classification system is the Terzaghi rock load classification, introduced over 45
years ago [2]. Since then this classification has been modified [3] and new rock mass classification
systems proposed. These systems took cognizance of the new advances in rock support technology,
namely rock bolts and shotcrete, as w.ell as addressing different engineering projects such as tunnels,
chambers, mines, slopes and foundations. Today there are many rock classification systems in
existence and the major ones are listed in Table 1. These classifications have been applied through-
out the world - in the USA [1-4], Canada [5], Europe [6-8] and elsewhere [9-14].
553
554 Rock Mass Classification

22.2 PURPOSES OF ROCK MASS CLASSIFICATION


Rock mass classifications were never intended as the ultimate solution to design problems, but
only as a means towards this end. In fact, some 20 years ago when work started on the major
classification systems now in use, the underground excavation scene worldwide was characterized by
haphazard site investigation programs and by limited, if any, design procedures. Rock mass
classifications were developed to create some order out of chaos in site investigation procedures and
to provide desperately needed design aids. They were not intended to replace analytical considera-
tions, field observations, or engineering judgment; they were simply to be design aids, forming only
a part of the rock engineer's 'bag of tools'. It is therefore appropriate to identify clearly the purpose
and benefits, as well as the pitfalls, of rock mass classifications before proceeding with a description
of any specific classification system.
Rock masses are classified for the following purposes: (i) to identify the most significant para-
meters influencing the behavior of a rock mass; (ii) to divide a particular rock mass formation into
a number of rock mass classes of varying quality; (iii) to provide a basis for understanding the
characteristics of each rock mass class; (iv) to derive quantitative data for engineering design; (v) to
recommend support guidelines for tunnels and mines; (vi) to provide a common basis for commun-
ication between engineers and geologists; and (vii) to relate the experience on rock conditions at one
site to the conditions and experience encountered at others.
The above items suggest three main benefits of rock mass classifications: (i) improving the quality
of site investigations by calling for the minimum input data as classification parameters; (ii)
providing quantitative information for design purposes; and (iii) enabling better engineering
judgment and more effective communication on a project.
When used correctly and for the purposes for which they were created, rock mass classifications
can be powerful aids in rock engineering. When abused, they can be counterproductive. The major
pitfalls are: (i) using rock mass classifications as the 'ultimate empirical cook book', i.e. ignoring
analytical and observational design methods; (ii) using one rock mass classification system only,
i.e. without cross-checking the results with at least one other system; (iii) using rock mass classifica-
tions without enough input data; and (iv) using rock mass classifications without full realization of
their conservative nature and their limits arising from the database on which they were developed.

22.3 CLASSIFICATION PARAMETERS


The output from any rock mass classification is only as good as the input data. It is therefore
important that the necessary basic input data are available and reliable.
Unlike other engineering materials, rock presents unique problems. First of all, rock is a complex
material varying widely in its properties, and in most rock engineering situations not one but
a number of rock types will be present. Furthermore, a choice of rock materials is only available if
there is a choice of alternative sites for a given project, although it may be possible, to some extent, to
reinforce the rock surrounding the excavation. Most of all, the rock engineer and geologist are
confronted with rock as an assemblage of blocks of rock material separated by various types of
discontinuities, such as joints, faults, bedding planes, etc. This assemblage constitutes a rock mass.
Consequently, the engineering properties of both intact rock and the rock mass must be considered.
The question immediately arises as to how the rock material is related to the rock mass. In
answering this question one must note, first of all, that the importance of the properties of intact rock
material will be generally overshadowed by the properties of the discontinuities in the rock masses.
However, this does not mean that the properties of the intact rock material should be disregarded
when considering the behavior of jointed rock masses. After all, if discontinuities are widely spaced
or if the intact rock is weak and altered, the properties of the intact rock may strongly influence the
gross behavior of the rock mass. Furthermore, a sample of a rock material sometimes represents
a small-scale model of the rock mass, since they have both gone through the same geological cycle.
Nevertheless, in general, the properties of the discontinuities are of greater importance than the
properties of the intact rock material.
The selection of the parameters of greatest significance for assuring rock mass stability is an
important issue. There appears to be no single parameter which can fully and quantitatively describe
a jointed rock mass for engineering purposes. Various parameters have different significance and
only taken together can they describe a rock mass satisfactorily.
The strength of the rock material is considered as the first necessary parameter because it
constitutes the strength limit of the rock mass. The uniaxial compressive strength of rock material
Classification of Rock Massesfor Engineering: The RMR System and Future Trends 555
can be determined in the field indirectly by means of index tests so that one is not restricted to
laboratory testing.
The second parameter most commonly employed is the rock quality designation (RQD). This is
a quantitative index [3] based on a modified core recovery procedure which incorporates only
sound pieces of core 100 mm or greater in length. The RQD is a measure of drill core quality or
fracture frequency, and disregards the influence of joint tightness, orientation, continuity and gouge
(infilling). Consequently, the RQD does not fully describe a rock mass.
Other essential parameters used in describing rock masses are: spacing of discontinuities; condi-
tion of discontinuities (roughness, continuity, separation, joint-wall weathering, infilling); orienta-
tion of discontinuities; groundwater conditions (inflow, pressure); and in situ stresses.
It is accepted that in the case of surface excavations and those near-surface tunnels which are
controlled by the structural geological features, the strength of the intact rock material, the spacing,
condition and orientation of the discontinuities and the groundwater conditions are important
parameters. In the case of mines or deep tunnels and chambers where the behavior of rock masses is
stress controlled, knowledge of the virgin stress field or the changes in stress can be of greater
significance than the geological discontinuities. Most civil engineering projects will fall into the first
category of geologically controlled rock mass structures.
In essence, the following parameters are considered to be most important for rock mass classifica-
tion [15]: (i) the strength of the rock material, which constitutes the upper strength limit of the rock
mass. Because of the prominence of the compressive stress in rock engineering problems, the uniaxial
compressive strength is the usual parameter; (ii) the rock quality designation (RQD), a quantitative
index widely used in rock engineering, affording comparisons of rock behavior in varied engineering
situations; (iii) the basic geological parameters usually included in any geological survey, such as
spacing, orientation and condition (i.e. roughness, separation, continuity, weathering and infilling)
of discontinuities; (iv) groundwater conditions; (v) stress field; and (vi) major faults and folds.
A thorough discussion of the methods for quantitative description of discontinuities in rock
masses will be found in Chapters 8-11 of this reference work, and in a document issued by the ISRM
[16].

22.4 MAJOR CLASSIFICATIONS CURRENTLY IN USE


Of the many rock mass classification systems in existence today, six should be mentioned because
they are important contributions, namely those proposed by Terzaghi [2], Lauffer [6], Deere [3],
Wickham et ale [4], Bieniawski [17] and Barton et ale [7].
The 1946 rock load classification of Terzaghi [2] was the first practical classification system
introduced and has been dominant in the USA for many years, proving very successful for tunneling
with steel supports. Lauffer's classification [6] was a considerable step forward in the art of
tunneling in 1958 since it introduced the concept of the stand-up time of an unsupported span in
a tunnel, which is highly relevant in determining the type and amount of tunnel support. Deere's
classification [3] of 1967 introduced the rock quality designation (RQD) index - a simple and
practical method of describing and classifying the quality of rock core recovered from boreholes.
The concept of rock structure rating (RSR), developed in the United States in 1972 by Wickham
et ale [4], was the first system featuring classification ratings for weighing the relative importance of
classification parameters. The RMR system (also called the Geomechanics Classification), proposed
by Bieniawski in 1973 [9], and the Q system proposed by Barton et ale [7] in 1974, also include
different parameter ratings but, in addition, both provide quantitative data for the selection of
modern tunnel reinforcement measures such as rock bolts and shotcrete. Moreover, in addition to
mining the RMR system has been applied to rock slopes and foundations and ground rippability
assessments [1].
Today, the RMR system and the Q system are the two rock mass classifications most commonly
used in rock engineering around the world.

22.5 THE ROCK MASS RATING SYSTEM


The Rock Mass Rating (RMR) system, otherwise known as the Geomechanics Classification, was
developed by the author [9] during 1972-73. It was modified over the years as more case histories
became available and to conform with international standards and procedures [17]. Over the past
20 years, the RMR system has stood the test of time and benefited from extensions and applications
556 Rock Mass Classification

by many authors throughout the world. These varied applications, amounting to 351 case histories
[1], involved tunnels, chambers, mines, slopes and foundations. Nevertheless, it is important that the
RMR system is used for the purpose for which it was developed, and not as a 'cookbook' for
empirical design.

22.5.1 Definition of the RMR System


Due to the RMR system having been modified several times, and since the method is interchange-
ably known as the Geomechanics Classification or the Rock Mass Rating system, it is important to
state that the system has remained essentially the same in principle in spite of the changes. Thus, any
modifications and extensions were the outgrowth of the same basic method and should not be
misconstrued as new systems. To avoid any confusion, Table 1 lists all the extensions of the RMR
system which were valuable new applications but still part of the same overall RMR system [10, 14,
18-22].
Furthermore, some users of the RMR system list their results as 'CSIR rating' or talk of the 'CSIR
Geomechanical' system. This is incorrect and it has never been used or suggested to this effect by the
author. The correct expressions are 'Rock Mass Rating system', 'RMR system' or the 'Geomechanics
Classification'. While it is true that the author has worked for an organization whose initials were
CSIR, that group did not develop the system and most of the work on his system was performed
after he left the CSIR over 15 years ago.

22.5.2 Classification Procedure


The following six parameters are used to classify a rock mass using the RMR system
(Geomechanics Classification).
1. Uniaxial compressive strength of rock material.
2. Rock quality designation (RQD).
3. Spacing of discontinuities.
4. Condition of discontinuities.

Table 1 Major Rock Mass Classifications Currently in Use [1]

N arne of classification Originator Country of Applications


and date origin

Rock loads Terzaghi, 1946 USA Tunnels with steel support


Stand-up time Lauffer, 1958 Austria Tunneling
NATM Rabcewicz, Pacher and Austria Tunneling
Miiller, 1964
Rock quality Deere, 1967 USA Core logging, tunneling
designation
RSR concept Wickham et al., 1972 USA Tunneling
RMR system Bieniawski, 1973 South Africa Tunnels, mines,
(Geomechanics slopes, foundations
Classification) (last modified, 1979, USA)
RMR system extensions Laubscher, 1977 South Africa Mining
Ghose and Raju, 1981 India Coal mining
Kendorski et al., 1983 USA Hard rock mining
Serafim and Pereira, 1983 Portugal Foundations
Gonzales de Vallejo, 1983 Spain Tunneling
Unal, 1983 USA Roof bolting/coal
Romana, 1985 Spain Slope stability
Newman, 1985 USA Coal mining
Venkateswarlu, 1986 India Coal mining
Robertson, 1988 Canada Slope stability
Q system Barton et al., 1974 Norway Tunnels, chambers
Strength-Size Franklin, 1975 Canada Tunneling
Basic geotechnical International Society for Rock General, communication
description Mechanics, 1981
Classification of Rock Masses for Engineering: The RMR System and Future Trends 557

5. Groundwater conditions.
6. Orientation of discontinuities.
To apply the RMR system, the rock mass is divided into a number of structural regions such that
certain features are more or less uniform within each region. Although rock masses are discontinu-
ous in nature, they may, nevertheless, be uniform in regions when, for example, the type of rock or
the discontinuity spacings are the same throughout that region. In most cases, the boundaries of
structural regions will coincide with major geological features such as faults, dykes, shear zones, etc.
After the structural regions have been identified, the classification parameters for each structural
region are determined from measurements in the field and entered onto the input data sheet shown
in Figure 1.
The RMR system is presented in Table 2. In section (a) of Table 2, five parameters are grouped
into five ranges of values. Since the various parameters are not equally important for the overall
classification of a rock mass, importance ratings are allocated to the different value ranges of the
parameters, a higher rating indicating better rock mass conditions. The importance ratings are
assigned to each parameter according to section (a) of Table 2. In this respect, the average typical
conditions are evaluated for each discontinuity set and the ratings are interpolated, using the
classification charts in Figures 2, 3, 4 and 5, and the guidelines in Table 3. The charts are helpful for
borderline cases and also remove the impression that abrupt changes in ratings occur between
categories. Figure 5 is used if either RQD or discontinuity data are lacking. Based on the correlation
data from Priest and Hudson [37], Figure 5 enables an estimate of the missing parameter.
Furthermore, it should be noted that the importance ratings given for discontinuity spacings apply
to rock masses having three sets of discontinuities. Thus, when only two sets of discontinuities are
present a conservative assessment is obtained. In this way, the number of discontinuity sets is
considered indirectly. Laubscher [10] presented a rating concept for discontinuity spacings as
a function of the number of joint sets. It can be shown that when less than three sets of
discontinuities are present, the rating for discontinuity spacing may be increased by 30%.

15
14
13
12
II
10
9
g' 8
~ 7
a:: 6
5
4
3
2
I

o 40 80 120 160 200 240


Uniaxia I compressive strength (MPa)

Figure 2 The RMR system: ratings for the strength of intact rock material

20
1.9
18
17
16
15
14
13
.=
0' ~2
I I
'5 10
a:: 9
8
7
6
5
4
3
2
I L...-_ _--'--_ _---'- ~ ......

o 20 40 60
RQD (0/0>

Figure 3 The RMR system: ratings for RQD


INPUT DATA FORM: GEOMECHANICS CLASSIFICATION (ROCK MASS RATING SYSTEM)
Name of the project:
Site of survey: STRUCTURAL DEPTH, m ROCK TYPE CONDITION OF DISCONTINUITIES VI
VI
Cond ucted by: REGION 00
Date: PERSISTENCE (CONTINUITY) Set 1 Set 2 Set 3 Set 4

STRENGTH OF INTACT ROCK MATERIAL DRILL CORE QUALITY R.Q.D. Very low: <1m
Low: 1-3 m
Uniaxial Point-load Medium: 3-10m
Designation compressive 0 R strength Excellent quality: 90-1 00% ........ High: 10-20 m
strength, M Pa index, MPa Good quality: 75-90% ........
Very high: >20m
Fair quality: 50-75% ........
Very high: Over 250. . . .. > 10 . Poor quality: 25-50% ........ SEPARATION (APERTURE)
High: 100-250 4-10 . Very poor quality: <25% ....... . Very tight joints: < 0.1 mm ......... . .
Medium High: 50-100 2-4 . Tight joints: 0.1-0.5 mm ......... . .
Moderate: 25-50 1-2 .
Moderately open joints: 0.5-2.5 mm .. . . . . . . . . .
Low: 5-25 <1 I R.Q.D. = Rock Quality Designation
Very Low: 1-5 .. Open joints: 2.5-10 mm ......... . .
Very wide aperture: > 10 mm ......... . .
STRIKE AND DIP ORIENTATIONS ROUGHNESS (state also if surfaces are stepped, undulating or planar)
Very rough surfaces: ......... ......... . .
Set 1 Strike............... (from. . . . . .. to ) Dip: .
(average) (angle) (direction) Rough surfaces: ......... ......... . .
Set 2 Strike............... (from. . . . . .. to ) Dip: . Slightly rough surfaces: ......... ......... . .
Set 3 Strike............... (from. . . . . .. to ) Dip: . Smooth surfaces: ......... ......... . .
Set 4 Strike............... (from. . . . . .. to ) Dip: . Slickensided surfaces: ......... . .
NOTE: Refer all directions to magnetic north. FILLING (GOUGE)
Type:
SPACING OF DISCONTINUITIES Thickness:
Uniaxial compressive strengt~, M Pa ......... ......... . .
Rock Mass Classification

Set 1 Set 2 Set 3 Set 4


Seepage:
Very wide: Over 2 m WALL ROCK OF DISCONTINUITIES
Wide 0.6-2 m Unweathered
Moderate: 200-600mm Slightly weathered
Close: 60-200mm Moderately weathered
Very close: <60mm Highly weathered
GROUND WATER Completely weathered
Residual soil
INFLOW per 10m liters/minute GENERAL CONDITIONS (completely dry,
GENERAL REMARKS AND ADDITIONAL DATA
of tunnel length damp, wet, dripping or flowing under
or low/medium or high pressure): MAJOR FAULTS specify locality, nature and orientations.
WATER PRESSURE kPa

IN SITU STRESSES NOTE: For definitions and methods consult ISRM document: ' Quantitative description of
discontinuities in rock masses'.

Figure 1 Input data form for engineering classification of rock masses


Classification of Rock Massesfor Engineering: The RMR System and Future Trends 559
20
19
18
17
16
15
14
13
12
g' I I
'B I~
0::: 8
7
6
5
4
3
2
I 10.0-_--'-_ _....10.--_---'_ _.........._ - - - - - '
o 400 800 1200 1600 2000
Spacing of discontinuities (mm)

Figure 4 The RMR system: ratings for discontinuity spacing

100
35 40

90
80 ROD max
70
60
~ Numbers in graph
0 50 represent combined ROD and
0
0: spacing ratings of each region
40
30 - - -- Average correlation line

20
10
0
10 20 30 40 60 100 200 600 2000
Mean discontinuity spacing (mm)

Figure 5 The RMR system: correlation between RQD and discontinuity spacing (after Priest and Hudson [37])

After the importance ratings of the classification parameters are established, the ratings for the five
parameters listed in section (a) of Table 2 are summed to yield the basic (unadjusted for discontinu-
ity orientations) Rock Mass Rating for the structural region under consideration. The next step is to
include the sixth parameter, namely the influence of the strike and dip orientation of discontinuities,
by adjusting the basic Rock Mass Rating according to section (b) of Table 2. This step is treated
separately because the influence of discontinuity orientations depends upon the engineering applica-
tions, e.g. tunnel, mine, slope or foundation. It will be noted that the 'value' of the parameter
'discontinuity orientation' is not given in quantitative terms but by qualitative descriptions such as
'favorable'. To facilitate a decision whether strike and dip orientations are favorable or not in
tunneling, reference should be made to Table 4, which was originally based on studies by Wickham
et ale [4], but was modified significantly once sufficient RMR case histories were accumulated. For
slopes and foundations, the reader is referred to appropriate papers [22, 23].
The parameter 'discontinuity orientation' reflects on the significance of the various discontinuity
sets present in a rock mass. The main set is usually designated as set 1 and it controls the stability
of an excavation, e.g. in tunneling it will be the set whose strike is parallel to the tunnel axis. The
summed-up ratings of the classification parameters for this discontinuity set will constitute the
overall Rock Mass Rating. On the other hand, in situations where no one discontinuity set is
dominant and of critical importance, or when estimating rock mass strength and deformability,the
ratings from each discontinuity set are averaged for the appropriate individual classification
parameter.
In the case of civil engineering projects, an adjustment for discontinuity orientations will generally
suffice. For mining applications or deep tunnels, other adjustments may be called for, such as the
stress at depth or a change in stress, and these have been discussed by Laubscher [10] and
Kendorski et ale [19]. The procedure for these adjustments is depicted in Figure 6.
VI
~
o
Table 2 The Rock Mass Rating System (Geomechanics Classification) [17]

(a) Classification Parameters and their Ratings

Parameter Ranges of values

Point load strength For this low range, uniaxial


Strength of index (MPa) > 10 4-10 2-4 1-2 compressive test is
intact rock preferred
material
Uniaxial compressive > 250 100-250 50-100 25-50 5-25 1-5 < 1
strength (MPa)
Rating 15 12 7 4 2 1 0
2 Drill core quality RQD (0/0) 90-100 75-90 50-75 25-50 < 25
Rating 20 17 13 8 3
3 Spacing of discontinuities >2m 0.6-2 m 200-600 mm 60-200 mm <6Omm
Rating 20 15 10 8 5
Slickensided surfaces
Very rough surfaces Slightly rough surfaces Slightly rough surfaces or Soft gouge> 5 mm thick
Not continuous Gouge < 5 mm thick or
4 Condition of discontinuities Separation < 1 mm Separation < 1 mm
No separation
Highly weathered wall or Separation > 5 mm
Rock Mass Classification

Unweathered wall rock Slightly weathered walls


Separation 1-5 mm Continuous
Continuous
Rating 30 25 20 10 0
5 Groundwater
Inflow per 10m
tunnel length None < 10 10-25 25-125 > 125
(L min -1) or or or or or
Joint water
pressure
Ratio 0 < 0.1 0.1-0.2 0.2-0.5 > 0.5
Major principal
stress
or or or or or
General conditions Completely dry Damp Wet Dripping Flowing
Rating 15 10 7 4 0
(b) Rating Adjustment for Discontinuity Orientations

Strike and dip orientations of


discontinuities Very favorable Favorable Fair Unfavorable Very unfavorable

Tunnels and mines 0 -2 -5 -10 -12


Ratings Foundations 0 -2 -7 -15 -25
Slopes 0 -5 -25 - 50 -60

(c) Rock Mass Classes Determined from Total Ratings

Rating 100-81 80-61 60-41 40-21 <20


Class I II III IV V
Description Very good rock Good rock Fair rock Poor rock Very poor rock

(d) Meaning of Rock Mass Classes

Class I II III IV V

Average stand-up time 20 y for 15 m span 1 y for 10m span 1 week for 5 m span 10 h for 2.5 m span 30 min for 1 m span
Cohesion of the rock mass (kPa) >400 300-400 200-300 100-200 < 100
Friction angle of the rock mass (deg) >45 35-45 25-35 15-25 < 15
Classification of Rock Massesfor Engineering: The RMR System and Future Trends

VI
0'\
~
562 Rock Mass Classification
Table 3 The RMR System: Guidelines for Classification of Discontinuity Conditions [1]

Parameter a Ratings

Discontinuity length <1m 1-3 m 3-10m 10-20 m >20m


(persistence/continuity)
6 4 2 1 0
Separation (aperture) None <0.1 mm 0.1-1.0 mm 1-5 mm >5mm
6 5 4 1 0
Roughness Very rough Rough Slightly rough Smooth Slickensided
6 5 3 1 0
Hard filling Soft filling
Infilling (gouge) None <5mm >5mm <5mm >5mm
6 4 2 2 0
Weathering Unweathered Slightly Moderately Highly Decomposed
weathered weathered weathered
6 5 3 1 0

a Some conditions are mutually exclusive. For example, if infilling is present, it is irrelevant what the roughness may be, since

its effect will be overshadowed by the influence of the gouge. In such cases, use Table 2 directly.

Strength of
intact rock
Blasting damage
Rating: 0-15 ~
adjustment,Ab
0.8-1.0

Discontinuity
Discontinuity
Orientation
I
density f adjustment

I
Insifu stress and
ROD: 0-20

....
Spacing:0-20 change of stress
adjustment
Rating: 0-40 As
I
I
Basic RMR
0-100 I 0.6-1.2

Discontinuity
condition Major faults and
Rating: 0-30 l-- fractures
5
0.7-1.0

Adjusted RMR
1
Groundwater
condition RMR xA b x ~s xS
Rating: 0-15 ~
maximum 0.5

J
I Support recommendations
I
Figure 6 Adjustments to the RMR system for mining applications

After the adjustment for discontinuity orientations, the rock mass is classified according to
section (c) of Table 2, which groups the final (adjusted) Rock Mass Rating (RMR) into five rock
mass classes, the full range of the possible RMR values varying from zero to 100. Note that the rock
mass classes are in groups of 20 ratings each. This concept of rating a rock mass out of a maximum
value of 100 has an advantage over an open-ended system in that it allows us to get the sense of
a relative quality, or the lack of it, of a given rock mass in terms of its maximum potential.
Next, section (d) of Table 2 gives the practical meaning of each rock mass class by relating it to
specific engineering problems. In the case of tunnels, chambers and mines, the output from the RMR
system is the stand-up time and the maximum stable rock span for a given Rock Mass Rating, as
shown in Figure 7.
Classification of Rock Massesfor Engineering: The RMR System and Future Trends 563

Table 4 Effect of Discontinuity Strike and Dip Orientations in Tunneling


(after Bieniawski [17])

Orientation of strike Dip


< 20° 20-45° 45-90°

Perpendicular to Fair Favorable Very favorable


tunnel axis: drive
with dip
Perpendicular to Fair Unfavorable Fair
tunnel axis: drive
against dip
Parallel to Fair Fair Very unfavorable
tunnel axis

I day I week I month I year


30
C
20
Immediate
collapse
~~
(>,\~O;

10 ~(>C:;40
8
~oc'+-
:§ 6
5 30
c 4
c
Q.
eft 3

80:: 2

Stand-up time (h)

Figure 7 Relationship between the stand-up time and span for various rock mass classes according to the RMR system:
output for tunneling and mining. The plotted data points represent roof falls studied: filled squares for mines, open squares for
tunnels. The contour lines are limits of applicability.

When mixed quality rock conditions are encountered at the excavated rock face, such as 'good
quality' and 'poor quality' being present in one exposed area, it is essential to identify the 'most
critical condition' for the assessment of the rock strata. This means that the geological features which
are most significant for stability purposes will have an overriding influence. For example, a fault or
a shear in a high quality rock face will playa dominant role, irrespective of the high rock material
strength in the surrounding strata.
It is recommended that when there are two or more clearly different zones in one rock face, one
approach to adopt is to obtain RMR values for each zone and then compute the overall weighted
value by the surface area corresponding to each zone in relation to the whole area, as well as by the
influence that each zone has on the stability of the whole excavation.
The RMR system provides guidelines for the selection of rock reinforcements for tunnels, in
accordance with Table 5. These guidelines depend on such factors as the depth below the surface
(in situ stress), tunnel size and shape, and the method of excavation. Note that the support measures
given in Table 5 represent the permanent and not the primary or temporary support. Approximate
support guidelines, as suggested by Hoek [38], are depicted in Figure 8. Both Table 5 and Figure 8
are applicable to rock masses excavated using conventional drilling and blasting procedures.
Most recently, Lauffer [6] presented a revised stand-up time diagram specifically for tunnel-
boring machine (TBM) excavation and superimposed it on the RMR diagram given in Figure 7. The
result is depicted in Figure 9, which is most useful because it demonstrates how the boundaries of
564 Rock Mass Classification

Table S RMR System Guidelines for Excavation and Support in Rock Tunnels (after Bieniawski [17])8.

Rock mass Excavation Rockbolts (20 mm Support Steel sets


class diameter, fully grouted) Shotcrete

Very good Full face, 3 m advance Generally no support required except for occasional spot bolting
rock, I
RMR: 81-100
Good rock, II Full face, 1.0-1.5 m Locally bolts in crown 50 mm in crown where None
RMR: 61-80 advance. Complete 3 m long, spaced required
support 20 m from 2.5 m with occasional
face wire mesh
Fair rock, III Top heading and bench, Systematic bolts 4 m 50-100 mm in crown None
RMR: 41-60 1.5-3 m advance in long, spaced 1.5-2 m and 30 mm in sides
top heading. Com- in crown and walls
mence support after with wire mesh in
each blast. Complete crown
support 10 m from
face
Poor rock, IV Top heading and bench, Systematic bolts 4-5 m 100-150 mm in crown Light to medium ribs
RMR: 21-40 1.0-1.5 m advance in long, spaced 1-1.5 m and 100 mm in sides spaced 1.5 m
top heading. Install in crown and walls where required
support concurrently with wire mesh
with excavation 10 m
from face
Very poor Multiple drifts. Systematic bolts 5-6 m 150-200 mm in crown, Medium to heavy
rock, V 0.5-1.5 m advance in long, spaced 1-1.5 m 150 mm in sides and ribs spaced 0.75 m
RMR: ::::; 20 top heading. Install in crown and walls 50mm on face with steel lagging
support concurrently with wire mesh. Bolt and fore-poling if
with excavation. invert required. Close
Shotcrete as soon as invert
possible after blasting

8Shape: horsehoe; width: 10m; vertical stress: 25 MPa; construction: drilling and blasting.

Tunneling quality index, Q

0.001 0.01 0.1 1.0 10 40 100 400 1000


oX: 0 0
e
(/) u
(/)
Structura lIy
~(/) 00 0.1 controlled 0.1

.~ failure
~ 0.2 0.2
'0
0
"0
c
o 0
:) -
~
.c
0.3 0.3
.a Q,)
~ &
c
Q)"C ~
> 'iii t; 0.4 0.4
'iii to..
Q,)
~~ >
.iii 0.5 0.5
e
0.0 (/)
E
a~~~I~~~
~
8£ a.
E 0.6 0.6
E 0 stress
:) u
E Not
.kpz
0.7 0.7
'x0 "0
'x0 ---.. - practical to
maintain stable
:E 'r: 0.8 0.8
openings Stress -

t
:::>
induced
0.9 failure 0.9
0
';Q: 1.0 1.0
0 10 20 30 40 50 60 70 80 90 100

I Very poor Poor Fair Good I Very good


Rock mass ratings, RMR

Figure 8 Approximate support guidelines (after Hoek [38])


Classification of Rock Masses for Engineering: The RMR System and Future Trends 565

Iweek Imonth
30~----~--'--;~--..,..._...,.--~;"--~-"'r----

20

AA
10 A
8
B
E 6
c: 5 c
&.en 4
'8 3
o
Q: 2 40
{o'\~~
~o~~ \
30 ~oc¥.
E \ TBM classes

10

Stand-up time (h)

Figure 9 Modified 1988 Lauffer diagram depicting boundaries of rock mass classes for TBM applications (after Lauffer [6])

RMR classes are shifted for TBM applications. Thus, an RMR adjustment can be made for
machine-excavated rock masses.
The support load can be determined from the RMR system as proposed by Unal [20]
P = [(100 - RMR)/100]yB (1)

where P is the support load (kN), B is the tunnel width (m) and y is the rock density (kg m -3).
It must be emphasized that for all applications such as those involving the selection of rock
reinforcement and determination of rock loads or rock mass strength and deformability, it is the
actual RMR value that must be used and not the rock mass class within which this RMR value falls.
In this way, the RMR system is very sensitive to individual parameters because within one rock mass
class, e.g. 'good rock', there is much difference between RMR = 80 and RMR = 61.
Finally, note that the ranges in Table 2 follow the recommendations of the International Society
of Rock Mechanics (ISRM) Commission on Standardization and Classification. The interested
reader is referred to an ISRM document entitled 'Suggested Methods for Quantitative Description
of Discontinuities in Rock Masses' [16].

22.5.3 Applications of the RMR System


The RMR system has found wide applications in various types of engineering projects such as
tunnels, slopes, foundations and mines. Most of the applications have been in the field of tunneling
[15].
This classification system has also been used widely in mining, particularly in the USA, India and
Australia. Initially, Laubscher [10] applied the RMR system in asbestos mines in Africa. More
recently, the RMR system was applied to coal and hard rock mining [13, 14, 19, 20, 24].
The RMR system is also applicable to slopes [22] and to rock foundations [23]. This is a useful
feature which can assist with the design of slopes near the tunnel portals as well as allow estimates of
the deformability of rock foundations for bridges and dams. Other special uses include applications
to assess rock rippability, cuttability and cavability [1].

22.5.3.1 Rock mass deformability


In the case of rock foundations, knowledge of the modulus of deformability of rock masses is of
prime importance. The RMR system has proved to be a useful method for estimating the in situ
deformability of rock masses [25]. As shown in Figure 10, the following correlation was obtained
EM = 2RMR - 100 (2)

where EM is the in situ modulus of deformation in GPa and RMR > 50.
I
566 Rock Mass Classification

- 70
e.
C)
+ Bieniawski 119781
• Serafin and Pereira (1983) "
- 60 o Stephens and Banks (1989) Ie
tJ + I
+

2RMR-IOO'li +++
C
0
~ 50
c

.'
EM'

.eQ)E 40
"C
0/,'
1/
'0 30
CI)
~
"3 ~+
"C
0 20 +++0
E E _IO IRMR -IOl/40 ++1+
~
M \ . ,..: *
,,0.,"+.•
10
.~

~ ___ ..- ~.c;.


" ++/
00•+
0 20 40 60 80 100

Geomechanics rock mass rating (RMR)

Figure 10 Correlation between the in situ modulus of deformation and Rock Mass Rating

Subsequently, Serafim and Pereira [21] provided many results in the range RMR < 50 and
proposed a new correlation
EM = 10(RMR-l0)/40 (3)

In the case of slopes, the output is given in section (d) of Table 2 as the cohesion and friction of the
rock mass. Romana [22] has applied the RMR system extensively for the determination of rock
slope stability.

22.5.3.2 Rock mass strength


Hoek and Brown [26] proposed a method for estimating rock mass strength which makes use of
the RMR classification. The criterion for rock mass strength is as follows

(4)

where CT 1 is the major principal stress at failure, (13 is the applied minor principal stress, CT c is the
uniaxial compressive strength of the rock material, and m and s are constants dependent upon the
properties of the rock and the extent to which it has been fractured by being subjected to CT 1 and CT 3.
For intact rock m = mh and this is determined from a fit of the above equation to triaxial test data
from laboratory specimens, taking s = 1 for rock material.
For rock masses, the constants m and s are related to the basic (unadjusted) RMR as follows [27].
For undisturbed rock masses (smooth-blasted or machine-bored excavations)
m miexp[(RMR - 1(0)/28] (5)

s exp[(RMR - 1(0)/9] (6)

For disturbed rock masses (slopes or blast-damaged excavations)


m miexp[(RMR - 1(0)/14] (7)

s exp[(RMR - 1(0)/6] (8)

The typical values of m and s for various rock types and corresponding to various RMR values are
listed in Table 6.
Yudhbir [40] studied a rock mass criterion of the form proposed by Bieniawski [41]

(9)
Table 6 Approximate Relationship Between Rock Mass Quality and Material Constants (after Hoek and Brown [27])

Empirical failure criterion Carbonate rocks with Lithified argillaceous rocks: Arenaceous rocks with Fine-grained polyminerallic Coarse-grained polyminerallic
0' 1 = 0' 3 + mO' c 0' 3 + sO' ~
J well-developed crystal mudstone, siltstone, shale strong crystals and poorly igneous crystalline rocks: igneous and metamorphic
0' 1 = major principal effective stress cleavage: and slate (normal to developed crystal cleavage: andesite, dolerite, diabase crystalline rocks:
0' 3 = minor principal effective stress dolomite, limestone and cleavage) sandstone and quartzite and rhyolite amphibolite, gabbro, gneiss,
0' c = uniaxial compressive strength marble granite, norite, quartz-diorite
of intact rock and
m and s are empirical constants

Intact rock samples:


laboratory size specimens free ma 7.00 10.00 15.00 17.00 25.00
from discontinuities. sa 1.00 1.00 1.00 1.00 1.00
RMR = 100, Q = 500 mb 7.00 10.00 15.00 17.00 25.00
Sb 1.00 1.00 1.00 1.00 1.00
Very good quality rock mass:
tightly interlocking, m 2.40 3.43 5.14 5.82 8.56
undisturbed rock with unweathered s 0.082 0.082 0.082 0.082 0.082
joints at 1-3 m. RMR = 85, m 4.10 5.85 8.78 9.95 14.63
Q = 100 s 0.189 0.189 0.189 0.189 0.189

Good quality rock mass:


fresh to slightly weathered m 0.575 0.821 1.231 1.395 2.052
rock, slightly disturbed with s 0.00293 0.00293 0.00293 0.00293 0.00293
joints ar 1-3 m. RMR = 65, m 2.006 2.865 4.298 4.871 7.163
Q = 10 s 0.0205 0.0205 0.0205 0.0205 0.0205

Fair quality rock mass:


several sets of moderately m 0.128 0.183 0.275 0.311 0.458
weathered joints spaced s 0.00009 0.00009 0.00009 0.00009 0.00009
at 0.3-1 m. RMR = 4, Q = 1 m 0.947 1.353 2.030 2.301 3.383
s 0.00198 0.00198 0.00198 0.00198 0.00198
Poor quality rock mass:
numerous weathered joints m 0.029 0.041 0.061 0.069 0.102
at 30-500 mm, some gouge. s 0.000003 0.000003 0.000003 0.000003 0.000003
Clean compacted waste rock. m 0.447 0.639 0.959 1.087 1.598
RMR = 23, Q = 0.1 s 0.00019 0.00019 0.00019 0.00019 0.00019

Very poor quality rock mass:


numerous heavily weathered m 0.007 0.010 0.015 0.017 0.025
joints spaced < 50 mm with s o.()()()()()() 1 O.()()()()()()1 O.()()()()()()1 O.()()()()()()1 O.()()()()()()1
gouge. Waste rock with fines. m 0.219 0.313 0.469 0.532 0.782
Classification of Rock Massesfor Engineering: The RMR System and Future Trends

RMR = 3, Q = 0.01 s 0.()()()()2 0.()()()()2 0.()()()()2 0.()()()()2 0.()()()()2


v.
0\
-First stated m and s values are for disturbed rock mass. bSecond stated m and s values are for undisturbed rock mass. -.J
568 Rock Mass Classification

where ex = 0.75 and A is a function of rock mass quality (note that A = 1 for intact rock), namely
A = exp[(RMR - 1(0)/14]

and B depends on rock type as follows: shale and limestone, B = 2; siltstone and mudstone, B = 3;
sandstone and quartzite, B = 4; and norite and granite, B = 5. For coal, ex = 0.6 and B = 4.
Ramamurthy [39] introduced the following relationship
(10)

where a 1 is the stre-ngth of the rock mass, a 3 is the confining stress and a em is the unconfined
compressive strength of the rock mass given by
O"cm = O"cexp[(RMR - 100)/18.75]

where a c is the uniaxial compressive strength of the rock material and


Bm = Bexp[(RMR - 100)/75.5]

B depends on the rock type, namely: shale and sandstone, B = 2.2; limestone, B = 2.4; basalt,
B = 2.6; marble, B = 2.8; and granite, B = 3.0. ex is the slope of a line through log( a cia 3) versus
log(al - a3)/a3, falling within the narrow range 0.75-0.85. It is normally assumed that ex = 0.8.
Moreno Tallon [29] developed a series of correlations between tunnel deformation, RMR and
time, based on a case history in Spain. Unal [20] proposed an 'integrated approach' to roof stability
assessment in coal mines by incorporating RMR with roof span, support pressure, time and
deformation. This is diagrammatically depicted in Figure 11. Finally, recent research by Nicholson
[30], incorporating the RMR system, proposed an empirical, nonlinear, stress-dependent, constitut-
ive relationship for rock masses.

22.5.3.3 Advantages and limitations


The RMR system is very simple to use and the classification parameters are easily obtained from
borehole data or underground mapping. This classification method is applicable and adaptable to

100-RMR
p = yB
100

Ground reaction
curve
10
E
5 15
c:
.2
0 20
E
~Q) 25
0

30

Figure 11 Integration of Rock Mass Rating with support characteristics and roof deformation in coal mines
(after Unal [20])
Classification of Rock Masses for Engineering: The RMR System and Future Trends 569

many different rock engineering situations and lends itself to use with theoretical concepts [20, 27,
29, 30]. The method is also suitable for knowledge-based expert systems. With the application of
fuzzy set methodology to the RMR system [31], the subjectiveness or fuzziness, inherent in
a classification system, can be considered and incorporated into an expert system.
However, the output from the RMR classification method tends to be rather conservative, which.
could lead to overdesign of support systems. This aspect is best overcome by monitoring rock
behavior during tunnel construction and adjusting rock classification predictions to local condi-
tions.
Finally, the RMR system - as with any other classification system - is not to be taken as
a substitute for engineering design. This classification is only a part of the empirical design approach,
which is one of the three main design approaches in rock engineering (empirical, observational and
analytical). It should be applied intelligently and used in conjunction with observational and
analytical methods to formulate an overall design rationale compatible with the design objectives
and site geology [42].
For the convenience of the user, a microcomputer program has been developed for the determina-
tion of the Rock Mass Rating and the resulting rock mass properties [1].

22.5.3.4 An example of use


The following is a typical example of finding rock mass quality by the RMR system.
Given: Rock mass in horizontal bedding, 152 m (500 ft) below surface.
Rock material: shale.
Uniaxial compressive strength: 50 MPa (5800 psi).
Strata conditions: separation < Imm; slightly weathered, slightly rough surfaces, no infill-
ing.
RQD = 60%. Three discontinuity sets.
Spacing of main discontinuities: 150 mm (6 in).
Groundwater conditions: damp.
I n situ stresses: horizontal stress = 2.5 x vertical stress.
Solution using the RMR system.
Rating due to uniaxial compressive strength 5
Rating due to RQD 12
Rating due to discontinuity spacing 7
Rating due to condition of discontinuities 17
Rating due to groundwater conditions 10
Rating due to strike and dip orientations -5
(Horizontal bedding = 'fair' orientation)
ROCK MASS RATING 46
Factor due to in situ stresses 0.9
FINAL ROCK MASS RATING, RMR 41
(Rock mass class III: fair rock)
Rock mass modulus of deformation 5.9 GPa
Cohesion of the rock mass 200 kPa
Friction of the rock mass 25°
Rock mass strength parameters for the Hoek-Brown criterion:
smooth excavation, m = 1.353 disturbed excavation, m = 0.183
s = 0.0014 s = 0.00005.

22.5.4 Database for the RMR System


The database used for the development of a rock mass classification may indicate the range of its
applicability. For example, the RMR system originally involved 49 case histories (which were
reanalyzed by Unal [20]), followed by 62 case histories added by Newman and Bieniawski [24] and
a further 78 tunneling and mining case histories collected between 1984 and 1989. To date, the RMR
570 Rock Mass Classification
system has been used in 351 case histories [1]. To demonstrate the RMR database ranges,
histograms are given in Figures 12, 13 and 14, depicting the ranges of the RMR values, spans of
excavations, and depths below the surface applicable to the case histories on the basis of which the
RMR system was developed.
It was also found that the system could be successfully used in rock formations not featured in the
original case histories [33]. At the same time, in some cases, the system did not provide realistic
results [32]. Nakao et ale [12] made a significant contribution by performing a statistical reconsider-
ation of the parameters for the RMR system to assess its applicability to Japanese geological
conditions. In total, 152 tunnel cases were studied. It was found that the results of the parameter-
rating analysis 'virtually agreed with those of the RMR concept'.

22.5.5 Correlations with Other Classification Systems


A correlation has been proposed between the RMR system and the Q index [17], and between the
RMR and RSR systems [1]. Based on 111 case histories analyzed for this purpose (involving 62
Scandinavian cases, 28 South African cases and 21 case histories from the USA, Canada, Australia

80

60

V)
S
0
U
'to-
0 40
~
CD
.J:)
E
:J
Z
20

o I.&-~..a.-~..a.----a...a.-. ..a.-.---I~---I~---I"""'----I"""'- ______


<20 21-30 31-40 41-5051-60 61-70 71-80 81-90 >91

RMR range (m)

Figure 12 Distribution of RMR values in the case histories studied (after Bieniawski [1])

120

100

80
o
o
't:J
60
'0
~
CD
.J:)
E 40
:J
Z

20

o L&---I..L---L.a..-..Jo..jL.-~L..-.....,L",I~""""'-""""-"""'-~
<3 3-4 4-5 5-7 7-10 10-15 15-20 20-25 >25

- Span range (m)

Figure 13 The range of spans encountered in the RMR case histories (after Bieniawski [1] )
Classification of Rock Massesfor Engineering: The RMR System and Future Trends 571
80

60

.n
I0
Co>
-0 40
~
CD
~
E
::J
Z
20

O.&A-_~-""'-""""-"""I.-----a..-"""""'..-.I~""""""'_""_""'_'-'-
o o
10
N
V
10
I
8
N
10
N
10 I
N
~ ~
Depth ranoe (m)

Figure 14 The range of depths in the RMR case histories (after Bieniawski [1])

ExceptiOnally E~onaIIY
poor

• Ngi case studies


1
>-
o G~hankol~ ~ud.s J
• Other case studies
e Indian case studies
]

;f

• I
20 •

o.........---..a-..--......- -......- - - - - - -......- -..........


0.001 0.01 0.1 1.0 10 100 1000
Rock mall quality (a)

Figure 15 Correlation between the RMR and the Q index (after Jeghwa [36])

and Europe), the results are plotted in Figure 15 from which the following relationship is found to be
applicable for tunnels
RMR = 91nQ + 44 (11)

For mining drifts, Abad et ale [3"] analyzed 187 coal mine roadways in Spain, arriving at the
correlation
RMR = 10.S1nQ + 42 (12)

Rutledge [11] determined from seven tunneling projects the following correlation
RSR = O.77RMR + 12.4 (13)
572 Rock Mass Classification
Moreno Tallon [29] confirmed the above relationships from four tunneling projects in Spain,
while Jethwa et ale [36] further substantiated equation (11) on the basis of 12 projects in India. His
results are also plotted in Figure 15.

22.6 FUTURE TRENDS


Three aspects may be identified for future developments involving rock mass classifications,
namely: (i) extensions of the systems for applications to underground nuclear repositories, which
would necessitate incorporation of thermal stress effects; (ii) greater integration of rock mass
classifications to analytical modeling by using them to estimate the strength and deformability of
rock masses; and (iii) systematic use of rock mass classifications for assessing rock mass quality
whenever in situ tests and monitoring during construction are performed. In this manner, valuable
case histories will be accumulated.
Most of all, there is a need for establishing a comprehensive design methodology for rock
engineering in which the useful role of rock mass classifications would be clearly designated.

22.7 SUMMARY AND CONCLUSIONS


The role of rock mass classifications in site characterization and project planning was critically
reviewed, and their advantages as well as common pitfalls were identified. In particular, the Rock
Mass Rating system (Geomechanics Classification) was described in detail together with the scope of
its applications and the database on which it was developed.
It is concluded that after two decades of extensive use, rock mass classifications can indeed serve
as reliable design aids in rock engineering. More specifically, the RMR system has established itself
as a versatile method for assessing rock mass conditions on engineering projects. The system has
withstood the test of time and is backed by 351 case histories, demonstrating its reliability and ease
of use. However, as an empirical design method, any rock mass classification should be applied
intelligently and used in conjunction with observational and analytical m~thods to formulate an
overall design rationale compatible with the design objectives and site geology.

22.8 REFERENCES
1. Bieniawski z. T. Engineering Rock Mass Classifications, p. 251. Wiley, New York (1989).
2. Terzaghi K. Rock defects and loads on tunnel support. In Rock Tunneling with Steel Supports (Edited by R. V. Proctor
and T. White), pp. 15-99. Commercial Shearing Co., Youngstown, OH (1946).
3. Deere D. U. and Deere D. W. The RQD index in practice. In Proc. Symp. Rock Classificationfor Engineering Purposes,
Cincinnati (Edited by L. Kirkaldie), pp. 91-101. ASTM Special Technical Publication 984, Philadelphia (1988).
4. Wickham G. E., Tiedemann H. R. and Skinner E. H. Support determination based on geologic predictions. In Proc.
Conf. Rapid Excavation and Tunneling pp. 43-64. AIME, New York (1972).
5. Franklin 1. A.and Dusseault M. B. Rock Engineering, p. 601. McGraw-Hill, Toronto (1989).
6. Lauffer H. Zur Gebirgsklassifizierung bei Frasvortrieben. Felsbau 6, 137-149 (1988).
7. Barton N., Lien R. and Lunde J. Engineering classification of rock masses for the design of tunnel support. Rock M echo
6, 183-236 (1974).
8. Brook N. and Dharmaratne P. G. R. Simplified rock mass rating system for mine tunnel support. Trans. Inst. Min.
Metall. 94, A148-A154 (1985).
9. Bieniawski Z. T. Engineering classification of jointed rock masses. Trans. S. Afr. Inst. Civil Eng. IS, 335-344 (1973).
10. Laubscher D. H. Geomechanics Classification of jointed rock masses - mining applications. Trans. Inst. Min. Metall.
86, AI-A7 (1977).
11. Rutledge 1. C. and Preston R. L. Experience with engineering classifications of rock. In Proc. Int. Symp. Tunneling,
Tokyo, pp. A3-A7. Japan Tunneling Society (1978).
12. Nakao K., Iihoshi S. and Koyama S. Statistical reconsiderations on the parameters for Geomechanics Classification, In
Proc. 5th Int. Congr. Rock Mech., Melbourne, pp. BI3-BI6. Balkema, Rotterdam (1983).
13. Ghose A. K. and Raju N. M. Characterization of rock mass vis-a-vis application of rock bolting in Indian coal
measures, In Proc. 22nd U.S. Symp. Rock Mech., Boston, MA (Edited by H. H. Einstein), pp. 422-427. MIT, Cam-
bridge, MA (1981).
14. Venkateswarlu V. Geomechanics Classification of Coal Measure Rocks vis-a-vis Roof Supports. Ph.D. Thesis, p. 251.
Indian School of Mines, Dhanbad (1986).
15. Bieniawski Z. T. Rock Mechanics Design in Mining and Tunneling, pp. 97-133. Balkema, Rotterdam (1984).
16. International Society for Rock Mechanics. ISRM Suggested Methods: Rock Characterization, Testing and Monitoring
(Edited by E. T. Brown), p. 211. Pergamon Press, Oxford (1982).
17. Bieniawski Z. T. The Geomechanics Classification in rock engineering applications, In Proc. 4th Int. Congr. Rock
Mech., Montreux, vol. 2, pp. 41-48. Balkema, Rotterdam (1979).
Classification of Rock Massesfor Engineering: The RMR System and Future Trends 573
18. Weaver J. Geological factors significant in the assessment of rippability. Civil Engineer in South Africa 17, 313-316
(1975).
19. Kendorski F., Cummings R., Bieniawski Z. T. and Skinner E. Rock mass classification for block caving mine drift
support. In Proc. 5th Int. Congr. Rock Mech., Melbourne, pp. B51-B63. Balkema, Rotterdam (1983).
20. Unal E. Design Guidelines and Roof Control Standards for Coal Mine Roofs. Ph.D. Thesis, p. 355. Pennsylvania State
University (1983).
21. Serafim J. L. and Pereira J. P. Considerations of the Geomechanics Classification of Bieniawski. In Proc. Int. Symp.
Engineering Geology and Underground Construction, Lisbon, vol 1, pp. 11.33-11.42. LNEC, Lisbon (1983).
22. Romana M. New adjustment ratings for application of Bieniawski classification to slopes. In Proc.. Int. Symp. Rock
Mechanics in Excavationsfor Mining and Civil Works, Mexico City, pp. 59-68. ISRM (1985).
23. Bieniawski Z. T. and Orr C. M. Rapid site appraisal for dam foundations by the Geomechanics Classification. In Proc.
12th Congr. on Large Dams, Mexico City, pp. 483-501. ICOLD (1976).
24. Newman D. A. and Bieniawski Z. T. Modified version of the Geomechanics Classification for entry design in under-
ground coal mines. Trans. Soc. Min. Eng. AIME 280, 2134-2138 (1986).
25. Bieniawski Z. T. Determining rock mass deformability: experience from case histories. Int. J. Rock Mech. Min. Sci.
& Geomech. Abstr. 15, 237-247 (1978).
26. Hoek E. and Brown E. T. Empirical strength criterion for rock masses. J. Geotech. Eng. Div., Am. Soc. Civ. Eng. 106,
1030-1035 (1980).
27. Hoek E. and Brown E. T. The Hoek-Brown failure criterion - a 1988 update. In Proc. 15th Can. Rock Mech. Symp.
University of Toronto (1988).
28. Stacey T. R. and Page C. H. Practical Handbook for Underground Rock Mechanics, p. 15. Trans Tech, ClausthaI,
Germany (1989).
29. Moreno Tallon E. Comparison and application of the Geomechanics Classification schemes in tunnel construction, In
Proc. Tunneling '82, pp. 241-246. Institution of Mining and Metallurgy, London (1982).
30. Nicholson G. A. and Bieniawski Z. T. An empirical constitutive relationship for rock mass. In Proc. 27th U.S. Symp.
Rock Mech. Tuscazoosa, AL (Edited by H. L. Hartman), pp. 760-766. AIME, New York (1986).
31. Nguyen V. U. and Ashworth E. Rock mass classification by fuzzy sets. In Proc. 26th U.S. Symp. Rock Mech., Rapid City,
SD (Edited by E. Ashworth), pp. 937-946. Balkema, Rotterdam (1985).
32. Kaiser P. K., MacKay C. and Gale A. D. Evaluation of rock classifications at B. C. Rail Tumbler Ridge tunnels, Rock
Mech. Rock Eng. 19, 205-234 (1986).
33. Singh R. N., Elmherig A. M. and Sunu M. Z. Application of rock mass characterization to the stability assessment and
blast design in hard rock surface,mining excavations. In Proc. 27th U.S. Symp. Rock Mech., Tuscazoosa, AL (Edited by
H. L. Hartman), pp. 471-478. AIME, New York (1986).
34. Goodman R. E. Introduction to Rock Mechanics, 2nd edn., pp. 42-49. Wiley, New York (1989).
35. Abad J., Celada B., Chacon E., Gutierrez V. and Hidalgo E. Application of Geomechanical Classification to predict the
convergence of coal mine galleries and to design their supports. In Proc. 5th Int. Congr. Rock Mech., vol. 2, Melbourne,
pp. E15-EI9. Balkema, Rotterdam (1983).
36. Jethwa J. L., Dube A. K., Singh B. and Mithal R. S. Evaluation of methods for tunnel support design in squeezing rock
conditions. In Proc. 4th Int. Congr. International Association of Engineering Geology, vol. 5, pp. 125-134. Balkema,
Rotterdam (1982).
37. Priest S. D. and Hudson J. A. Discontinuity spacings in rock. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr. 13,
135-148 (1976).
38. Hoek E. Geotechnical design of large openings at depth. In Proc. Conf. Rapid Excavation and Tunneling, pp. 1167-1180.
AIME, New York (1981).
39. Ramamurthy T. Stability of rock mass. Indian Geotech. J. 16, 1-74 (1986).
40. Yudhbir E. T. An empirical failure criterion for rock masses. In Proc. 5th Int. Congr. Rock Mech., vol. 1, Melbourne,
pp. BI-B8. Balkema, Rotterdam (1983).
41. Bieniawski Z. T. Estimating the strength of rock materials. J. S. Afr. Inst. Min. Metall. 74, 312-320 (1974).
42. Bieniawski Z. T. Design Methodology in Rock Engineering. Balkema, Rotterdam (1992).

You might also like