You are on page 1of 196

Chapter 1 Introduction 1

ROCK MECHANICS
FOR CIVIL ENGINEERS

Jian ZHAO

Professor of Rock Mechanics and Tunnelling


School of Architecture, Civil and Environmental Engineering
Swiss Federal Institute of Technology
Lausanne, Switzerland

AUTHOR’S NOTE

This current text is a preliminary draft of lecture notes under preparation


for the course of Rock Mechanics in the second year teaching for civil
engineering students.

Use the notes carefully as they may contain errors and mistakes.

Please let me know when you find any errors and mistakes.
Chapter 1 Introduction 2

Chapter 1
Introduction

1.1 Rock Mechanics as a Discipline

Rock mechanics is a discipline that uses the principles of mechanics to describe the
behaviour of rocks.

Here, the term of rock is in the scale of engineering. The scale is generally in the order of
between a few metres to a few thousand metres. Therefore, the rock considered in rock
mechanics is in fact the rock mass, which composes intact rock materials and rock
discontinuities.

What is so special of rock mechanics?

For normal construction materials, e.g., steel and concrete, the mechanical behaviours are
continuous, homogeneous, isotropic, and linearly elastic (CHILE). Properties of the man-
made materials are known and can often be controlled.

For rocks, due to the existence of discontinuities, the behaviours are discontinuous,
inhomogeneous, anisotropic, and non-linearly elastic (DIANE). Properties of the natural
geomaterials are unknown and often can not be controlled.

It is important to be award that in rock mechanics, rock discontinuities dominate the


mechanical and engineering behaviours. The existence of discontinuity depends on the
scale. The discontinuous nature and scale dependence feature is not common in other
man-made materials.

Rock mechanics is applied to various engineering disciplines: civil, mining, hydropower,


petroleum. In civil engineering, it involves foundation, slope and tunnel.

In structural engineering, the design process generally is as following:


Calculate external loading imposed on the structure;
Design the structure and analyse loading in structure elements;
Design the structure element and select materials.

In rock engineering, or geotechnical engineering, the whole process is different. Loading


condition is not easily calculateable, rock engineering, being sloping cutting or
underground excavation, does not impose loading, but disturbs the existing stress field of
the ground and redistribute the load. Therefore, the key process in rock engineering is to
understand the how the stress field is disturbed by engineering activities and how the rock
is behaving (responding) to the change of boundary conditions, and yet the material does
not has a characteristics controlled by man.

The objectives of learning rock mechanics are:


Chapter 1 Introduction 3

• To understand of the mechanical behaviour of rock materials, rock discontinuities and


rock masses.
• To be able to analyse and to determine mechanical and engineering properties of rocks
for engineering applications.

1.2 Use of This Book

This textbook is used as an introduction to rock mechanics at university engineering


teaching, with an emphasis to civil engineering. It is written as a self-contained book,
with a brief introduction of rock as a geological material and a review of basic mechanics
applicable to rock and rock discontinuity. The teaching of Chapters 2 and 3 can be
selective depending on the level of knowledge that students already have.

In the Swiss Federal Institute of Technology Lausanne (EPFL), rock mechanics class is a
28 hours course in the second year after learning already engineering geology and
mechanics, the teaching therefore focuses on Chapters 4, 5, 6 and 7. Subsequently, rock
engineering applications are taught at later years in EPFL. Depending on the teaching
hour available and overall teaching plan, the rock mechanics course could include
engineering applications in Chapters 8, 9, and 10.

Exercises are included in this textbook. It is highly recommended to teach this subject
with exercises.

At end of each chapters, recommended further reading materials are given. The
references cited in the text are listed at end of the book.
Chapter 2 Rock Formation and Rock Mass 1

CHAPTER 2
ROCK FORMATION AND ROCK MASS
This Chapter is intended as a review on the geological aspects of rock. Reader who is not
familiar with geology should start with one of the many textbooks on physical geology or
engineering geology, listed at the end of this Chapter.

2.1 Rock Formations and Types

2.1.1 Rock Formation and Rock Cycle

Rock is a natural geomaterial. In geological term, rock is a solid substance composed of


minerals, of which can consist in particulate form (soil particles) or in large form
(mountains, tectonic plates, planetary cores, planets). In common term, rock usually
stands a solid block of natural earth material.

Rocks are formed by three main origins: igneous rocks from magma, sedimentary rock
from sediments lithfication and metamorphic rocks through metamorphism.

Figure 2.1.1a shows the geological process involved in the formations of various rocks. It
should be noted that the processes are dynamic and continuous.

Figure 2.1.1a Rock cycle illustrating the role of various geological processes in rock
formation.

2.1.2 Rock Forming Minerals

All rocks are composed of minerals. A few rocks are composed by single minerals, but
most are by a group of minerals.

A mineral is a naturally occurring, inorganic solid with a definite chemical composition


and a crystalline structure. Minerals that are commonly found in rocks are often called
rock-forming minerals.

Silicate minerals form the largest group of minerals, and most rocks contain more than
95% of silicates. Silicates are composed largely of silicon and oxygen, with the addition
of ions such as aluminium, magnesium, iron, and calcium. Some important rock-forming
silicates include the feldspars, quartz, olivines, pyroxenes, amphiboles, garnets, and
micas.

A mineral can be identified by several physical properties. The key properties that are
important to influence rock properties are crystal structure, hardness and cleavage. Other
common properties include lustre, colour, fracture, and specific gravity. Table 2.1.2a
gives main properties of common rock forming minerals.
Chapter 2 Rock Formation and Rock Mass 2

Table 2.1.2a Principal properties of common monierals


Mineral Crystal form Cleavage Colour Hardness Specific
Gravity
Olivine Orthorhombic. 8-sided Cleavage rare seen, None when 6.5 3.2 - 3.6
and with 8 dome and irregular cracks fresh.
pyramid faces. common.

Pyroxene Short 8-sided, 2 cleavages at nearly Commonly 5-6 3.2 - 3.6


(Augite) prismatic crystals. 90°, not always well greenish to
developed. black.
Amphibole Long 6-sided prismatic 2 good cleavages Black to 5-6 2.9 - 3.2
(Hornblende) crystal. meeting at angles of light green.
54° and 124°.
Muscovite Thin, scale-like crystal Perfect 1 cleavage, Colourless, 2-3 2.8 - 3.1
and scaly yielding very thin, light colour
flexible scales in thick
pieces.
Biotite Thin, scale-like crystal, Perfect 1 cleavage, Black to 2.5 - 3 2.7 - 3.2
commonly 6-sided and yielding very thin, dark brown
scaly. flexible scales
Orthoclase Boxlike crystal. 1 perfect and I good White, grey, 6 2.5 - 2.6
feldspar Massive. cleavage at angle of pink, or
90°. pale yellow
Plagioclase Well-formed crystals 2 good cleavage at White or 6 - 6.5 2.6 - 2.7
feldspar and in cleavable or angle of 86°. grey, may
granular masses. be other
colour
Quartz 6-sided prismatic None or very poor Colourless 7 2.65
crystal, terminated by cleavage. or white,
6-sided triangular may be any
faces. colour

A crystal structure is the orderly geometric spatial arrangement of atoms in the internal
structure of a mineral. The crystal structure is based on regular internal atomic or ionic
arrangement that is often expressed in the geometric form that the crystal takes. A
mineral may show good crystal form (Figure 2.1.2a). However, in rock, they are often
massive, granular or compact with only microscopically visible (Figure 2.1.2b).

Figure 2.1.2a A fully developed quartz crystal.

Figure 2.1.2b Quartz minerals in a granite do not have visible crystal forms.

Crystal structure greatly influences a mineral's physical properties. For example, though
diamond and graphite have the same composition (both are pure carbon), graphite is very
soft, while diamond is the hardest of all known minerals. This happens because the carbon
atoms in graphite are arranged into sheets which can slide easily past each other, while the
carbon atoms in diamond form a strong, interlocking three-dimensional network.
Chapter 2 Rock Formation and Rock Mass 3

Minerals have different hardness, due to different crystal structures. The physical
hardness of a mineral is usually measured according to the Mohs scale. This scale is
relative and starts with 1 being the softest to 10 being the hardest. The minerals that
define the scale are: 1 – talc, 2 – gypsum, 3 – calcite, 4 – fluorite, 5 – apatite, 6 –
orthoclase feldspar, 7 – quartz, 8 – topaz, 9 – corundum, and 10 – diamond.

Cleavage is the tendency of crystalline materials to split along definite planes, creating
smooth surfaces. Some minerals have well developed cleavages. For example, mica
minerals have one perfect cleavage, and calcite has three perfect cleavages.

A rock is an aggregate of one or more minerals. Some rocks are predominantly composed
of just one mineral. For example, limestone is composed almost entirely of the mineral
calcite. Other rocks contain many minerals. For example, granite is composed of quartz,
feldspars and mica minerals. The specific minerals in a rock can vary widely. Some
minerals, like quartz, mica or feldspar are common, while others have been found in very
limited locations.

As rocks are composed of minerals, the properties of minerals become part of the
properties of rocks. However, the physical and mechanical properties of a rock are
governed by not only the intrinsic properties of minerals forming the rock, but also the
texture structures of the composition minerals in that rock.

2.1.3 Igneous Rocks

Igneous rocks are formed when molten rock (magma) cools and solidifies, with or without
crystallization. They can be formed below the surface as intrusive (plutonic) rocks, or on
the surface as extrusive (volcanic) rocks. This magma can be derived from either the
Earth's mantle or pre-existing rocks made molten by extreme temperature and pressure
changes. Figure 2.1.1a shows the origin of magma and igneous rock through the rock
cycle.

As magma cools, minerals crystallize from the melt at different temperatures. The magma
from which the minerals crystallize is rich in only silicon, oxygen, aluminium, sodium,
potassium, calcium, iron, and magnesium minerals. These are the elements which
combine to form the silicate minerals, which account for over 90% of all igneous rocks.

Igneous rocks make up approximately 95% of the upper part of the Earth's crust, but their
great abundance is hidden on the Earth's surface by a relatively thin but widespread layer
of sedimentary and metamorphic rocks.

Intrusive igneous rocks are formed from magma that cools and solidifies within the earth.
Surrounded by pre-existing rock (country rock), the magma cools slowly, and as a result
these rocks are coarse grained. The mineral grains in such rocks can generally be
identified with the naked eye. Intrusive rocks can also be classified according to the shape
Chapter 2 Rock Formation and Rock Mass 4

and size of the intrusive body and its relation to the other formations into which it
intrudes. Typical intrusive formations are batholiths, stocks, laccoliths, sills and dikes.

Extrusive igneous rocks are formed at the Earth's surface, when magma rises and reaches
the surface, either beneath water or air. The magma is then called lava. Because lava cools
and crystallizes rapidly, it is fine grained. If the cooling has been so rapid as to prevent
the formation of even small crystals the resulting rock may be a glass (such as the rock
obsidian).

Because of this fine grained texture it is much more difficult to distinguish between the
different types of extrusive igneous rocks than between different types of intrusive
igneous rocks. Generally, the mineral constituents of fine grained extrusive igneous rocks
can only be determined by examination of thin sections of the rock under a microscope,
so only an approximate classification can usually be made in the field.

Igneous rocks are classified according to mode of occurrence, texture, chemical


composition, and the geometry of the igneous body. In a simplified classification, igneous
rock types are separated on the basis of mineral composition and grain size, as shown in
Table 2.1.3a.

Table 2.1.3a Classification of common igneous rocks


Granitic Andesitic Basaltic Ultramafic
(acid) (felsic) (intermediate) (basic) (mafic) (ultrabasic)
Intrusive
(coarse grain) Granite Diorite Gabbro Peridotite

Extrusive
(fine grain) Rhyolite Andesite Basalt None

Silica Content >65% Silica 50-65% Silica 40-50% Silica <40% Silica
Quartz Amphibole
Main Mineral Orthoclase Plagioclase Ca-Plagioclase Olivine
Composition N-Plagioclase Biotite Pyroxene Pyroxene

Muscovite
Minor Mineral Biotite Pyroxene Olivine Ca-Plagioclase
Composition Amphibole Amphibole

Colour Light Dark

Igneous rocks which have crystals large enough to be seen by the naked eye are called
phaneritic; those with crystals too small to be seen are called aphanitic. Generally
speaking, phaneritic implies an intrusive origin; aphanitic an extrusive one. The crystals
embedded in fine grained igneous rocks are termed porphyritic. The porphyritic texture
develops when some of the crystals grow to considerable size before the main mass of the
magma consolidates into the finer grained uniform material. Some of the typical igneous
rocks are shown in Figure 2.1.3a.
Chapter 2 Rock Formation and Rock Mass 5

Figure 2.1.3a Igneous rocks, (i) granite, (ii) basalt, and (iii) prophyritic andesite.

2.1.4 Sedimentary Rocks

Sedimentary rock is formed in three main ways – by the deposition of the weathered
remains of other rocks (known as 'clastic' sedimentary rocks); by the deposition of the
results of biogenic activity; and by precipitation from solution. Sedimentary rocks include
common types such as sandstone, conglomerate, clay, shale, chalk and limestone.
Sedimentary rocks cover 75% of the Earth's surface, but count for only 5% of the rock in
the earth crust. Four basic processes are involved in the formation of a clastic sedimentary
rock: weathering (erosion), transportation, deposition and compaction, as sown in
Figure 2.1.1.

All rocks disintegrate slowly as a result of mechanical weathering and chemical


weathering. Mechanical weathering is the breakdown of rock into particles without
producing changes in the chemical composition of the minerals in the rock. Chemical
weathering is the breakdown of rock by chemical reaction.

Rock particles of different sizes are transported by the agents of erosion (usually water,
and less frequently by ice and wind). These agents reduce the size of the particles, often
sort them by size, and then deposit them in new locations. The sediments are usually
deposited in layers, generally at a lower elevation. As sediment deposition builds up, the
overburden pressure squeezes the sediment into layered solids in a process known as
lithification.

Clastic sedimentary rocks are composed of discrete fragments or clasts of materials


derived from other rocks. They are composed largely of quartz with other common
minerals including feldspars, amphiboles, clay minerals, and other minerals. Clastic
sedimentary rocks are commonly classified by their grain size, with shale and clay being
the finest with particles less than 0.004 mm, siltstone at 0.004 to 0.06 mm, and sandstone
at 0.06 to 0.2 mm, and conglomerates and breccias being the coarsest with grains 2 to 256
mm. Composition of the particles, the cement, and the matrix also contribute to the
classification. Classification of sediments and clastic rocks are shown in Table 2.1.4a,
while some typical clastic sedimentary rocks are shown in Figure 2.1.4a.

Table 2.1.4a Classification of common clastic sedimentary rocks

Particle size Comments Rock name


> 2 mm Rounded rock fragment Conglomerate
Angular rock fragment Breccia
1/16 - 2 mm Quartz with other minerals Sandstone
< 1/16 mm Split into thin layers Shale
Break into clumps or blocks Mudstone
Chapter 2 Rock Formation and Rock Mass 6

Figure 2.1.4a Sedimentary rocks, (i) shale, (ii) sandstone, (iii) conglomerates, (iv) rock
salt.

Texture of chemical sedimentary rocks is generally non-clastic. Main chemical


sedimentary rocks are limestone, dolomite, halite (salt) and gypsum.

2.1.5 Metamorphic Rocks

Metamorphic rock is a new rock type transformed from an existing rock type, through
metamorphism. When an existing rock is subjected to heat and extreme pressure, the rock
undergoes profound physical and/or chemical change. The existing rock may be
sedimentary rock, igneous rock or another older metamorphic rock (Figure 2.1.1a).

Metamorphic rocks make up a large part of the Earth's crust and are classified by texture
and by mineral assembly. They are formed deep beneath the Earth's surface by great
stresses from rocks above and high pressures and temperatures, known as regional
metamorphism. The high temperatures and pressures in the depths of the Earth are the
cause of the changes, and if the metamorphosed rocks are uplifted and exposed by
erosion, they may occur over vast areas at the surface. Metamorphic rocks are also formed
by the intrusion of molten rock (magma) into solid rock and form particularly at the place
of contact between the magma and solid rock where the temperatures are high, known as
contact metamorphism. Recrystallization of the rock will destroy the textures and fossils
present in sedimentary rocks.

Another important mechanism of metamorphism is that of chemical reactions that occur


between minerals without them melting. In the process atoms are exchanged between the
minerals, and thus new minerals are formed. Many complex high-temperature reactions
may take place, and each mineral assemblage produced provides us with a clue as to the
temperatures and pressures at the time of metamorphism.

Heat and pressure are the causes of metamorphism. When above 200°C, heat causes
minerals to recrystallise. Pressure forces some crystals to re-orient. The combined effects
of recrystallisation and re-orientation usually lead to foliation, which is an unique feature
of metamorphic rocks. It occurs when a strong compressive force is applied from one
direction to a recrystallizing rock. This causes the platy or elongated crystals of minerals,
such as mica and chlorite, to grow with their long axes perpendicular to the direction of
the force. This results in a banded, or foliated, rock, with the bands showing the colours
of the minerals that formed them.

Textures of metamorphic rocks are separated into foliated and non-foliated categories.
Foliated rock is a product of differential stress that deforms the rock in one plane,
sometimes creating a plane of cleavage: for example, slate is a foliated metamorphic rock,
originating from shale. Non-foliated rock does not have planar patterns of stress.
Chapter 2 Rock Formation and Rock Mass 7

Rocks that were subjected to uniform pressure from all sides, or those which lack
minerals with distinctive growth habits, will not be foliated. The common metamorphic
rocks are: slate (very fine-grained, slaty cleavage foliation); phyllite (fine-grained, slaty
cleavage foliation); schist (coarse-grained, schistocity foliation); gneiss (very coarse-
grained, gneisocity or banded foliation); marble (non-foliated, originating from
limestone); and quartzite (non-foliated, originating from quartz sandstone). The common
metamorphic rocks are shown in Table 2.1.5a and Figure 2.1.5a.

Table 2.1.5a Classification of common metamorphic rocks


Rock Texture Metamorphic grade Original parent rock
Slate Foliated Low grade Shale (clay minerals)
Phyllite Foliated Low to intermediate grade Shale
Mica schist Foliated Low to intermediate grade Shale
Chlorite schist Foliated Low grade Basalt
Gneiss Foliated High grade Granite, shale, andesite
Marble Non-foliated Low to high grade Limestone, dolomite
Quartzite Non-foliated Intermediate to high grade Quartz sandstone

Figure 2.1.5a Metamorphic rocks, (i) slate, (ii) schist, (iii) gneiss, and (iv) quartzite.

2.1.6 Rock Texture

Rock texture usually indicates the origin of rocks, how they formed, and their appearance.
Sedimentary, igneous and metamorphic rocks have different textures due to their different
origin. The two main texture forms are clastic and interlocking.

Clastic texture is typically found in clastic sedimentary rocks (Figure 2.1.6a). Description
of clastic texture includes shape and size, composition, and matrix. Clastic shape includes
form and rounding. Form indicates whether a grain is more spherical or platy. Roundness
refers to the degree of sharpness of the corners and edges of a grain. Composition
indicates the derivation of a rock's sediments. For instance, volcanic fragments, well-
rounded sands and cornered gravel all imply different sources. Matrix and cementation
indicates how a sedimentary rock holding together.

Figure 2.1.6a Clastic textures of sanstones.

Igneous and metamorphic rocks generally have microstructures that grains are
interlocked. During cooling, minerals are crystallised, and each requires space for its
crystal growth. It results in mineral crystals penetrate into other minerals, and form the
interlocking structure, as shown in Figure 2.1.6b.
Chapter 2 Rock Formation and Rock Mass 8

Figure 2.1.6b Interlocking textures of some igneous and metamorphic rocks.

An igneous or metamorphic rock may have grains of more or less uniformly coarse or fine
sizes or have grains of different sizes (porphyritic), depending on the formation
environment. Igneous microstructure is a combination of cooling rate, nucleation rate,
eruption (if a lava), and magma composition. Metamorphic microstructure is influenced
by rock origin, heating conditions and duration and mineral growth.

Rock material strength is essentially a structural strength of the composition of the


minerals in a rock material. It is governed by the strength of the minerals as well as the
structural bonding (integration) of the minerals. The interlocking microstructures of
igneous and metamorphic rocks lead to generally high strength of rock material, while the
clastic microstructures of sedimentary rocks often lead to low rock material strength,
particularly when cementation is weak.

Any existing weakness, e.g., microcracks, pores, and weak mineral particles, within a
rock material matrix also has great influence on the strength of the rock material. When
the rock is subjected to a stress, the weak points start to fail (cracking) and cracks then
propagate leading to overall failure of the rock. Rocks often contain existing weak
microstructures, hence lead to large variation of strengths. For example, two granites may
have variations in mica, a soft mineral, and grain structures, which in turn, will have
different strength. When a rock is weathered, even to a very small degree, part of the
structure bonding is weakened and lead to substantial reduction in strength.

2.2 Rock Discontinuities

2.2.1 Joints

A geological joint is a generally planar fracture formed in a rock as a result of extensional


stress. Joints are always in sets. Joints do not have any significant offset of strata either
vertically or horizontally (Figure 2.2.1a).

Figure 2.2.1a Typical joints seen (i) one dominant set, (ii) three sets.

Joints can be formed due to erosion of the overlying strata exposed at the surface. The
removal of overlying rock results in change of stresses, and hence leads to the fracturing
of underlying rock. Joints can also be caused by cooling of hot rock masses, which form
cooling joints, as illustrated in Figure 2.2.1b. Columnar jointing or columnar basalts are
typical joint features by cooling. Joints are also formed by tectonic movement, for
example, by folding.
Chapter 2 Rock Formation and Rock Mass 9

Figure 2.2.1b Joint formation due to cooling and contraction of pluton.

Joints are often in sets. A joint set is a group of parallel joints. Typically, a rock mass
can have between one to a few joint sets.

Joints are the most common type of rock discontinuities. They are generally considered
as part of the rock mass, as the spacing of joints usually is between a few centimetres and
a few metres. Therefore, for an engineering project, joints are a constant feature of the
rock mass at site within the project scale.

2.2.2 Faults

Geologic faults are planar rock fractures which show evidence of relative movement.
Large faults within the Earth's crust are the result of shear motion and active fault zones
are the causal locations of most earthquakes. Earthquakes are caused by energy release
during rapid slippage along faults. The largest examples are at tectonic plate boundaries,
but many faults occur far from active plate boundaries. Since faults usually do not consist
of a single, clean fracture, the term fault zone is used when referring to the zone of
complex deformation associated with the fault plane. Figure 2.2.2a.

Figure 2.2.2a Faults, fault zone and shear zone.

A shear zone is a wide zone of distributed shearing in rock. Typically this is a type of
fault but it may be difficult to place a distinct fault plane into the shear zone. Shear zones
can be only inches wide, or up to several kilometres wide.

As faults, particularly fault zone and shear zone, are large scale geological features. They
are often dealt separately from the rock mass. Small scale single faults often have the
similar effects as a joint. The behaviour large scale fault and shear zones require specific
investigation and analysis, if a project is to be constructed over or close such zones.

2.2.3 Folds

The term fold is used in geology when originally flat and planar rock strata are bent as a
result of tectonic force or movement. Folds form under very varied conditions of stress.
Folds can be commonly observed in sedimentary formation and as well as in metamorphic
rocks (Figure 2.2.3a).

Figure 2.2.3a Folds in a sedimentary formation.


Chapter 2 Rock Formation and Rock Mass 10

Folds are usually not considered as part of the rock mass. However, folds can be of the
similar scale as the engineering project and hence the significance of folds on the
behaviour of the rock mass must be taken into consideration. It should be noted that fold
has huge variation of features.

Folds, particularly intense folds, are often associated with high degree of fracturing and
relatively weak and soft rocks. Although the folding feature may not be directly taking
into account of rock mass, but the results of folding is often reflected in the rock mass
consideration.

2.2.4 Bedding Planes

As sedimentary rocks are formed in layers, the interfaces between layers are termed as
bedding planes. Bedding plane therefore is a discontinuity separating different rocks
(Figure 2.2.4a). Bedding plane often can be fully closed and cemented.

Figure 2.2.4a Some typical bedding planes.

Bedding planes are isolated geological features to engineering activities. It mainly creates
an interface of two rock materials. However, some bedding planes could also become
potential weathered zones and pocket of groundwater. For example, an interface between
porous sandstone and limestone may lead to extensive weathering of the limestone, which
leads to cavities along the interface.

2.3 Rock Material and Rock Masses

2.3.1 Engineering Scale and Rock

Engineering in and on rock has different scales, varying from a few centimetres to a few
kilometres. A borehole can be typically around 8 cm while a mine can spread up to a few
km. For civil engineering works, e.g., foundations, slopes and tunnels, the scale of
projects is usually a few ten metres to a few hundreds metres (Figure 2.3.1a).

Figure 2.3.1a Scale of rock engineering.

When such engineering scale is considered, then rock in such scale is generally a mass of
rock at the site. This mass of rock, often termed as rock mass, is the whole body of the
rock in situ, consists of rock blocks and fractures, typically seen in Figure 2.3.1b.

Figure 2.3.1b Typical rock masses.


Chapter 2 Rock Formation and Rock Mass 11

2.3.2 Composition of Rock Mass

As shown in Figure 2.3.2, a rock mass contains (a) rock material, in the form of intact
rock plates, blocks and wedges, of various sizes, and (b) rock discontinuities that cuts
through the rock, in the forms of fractures, joints, and faults.

Rock masses = Rock materials + Rock discontinuities

In addition, rock mass may also include filling materials in the discontinuities and dyke
and sill igneous intrusions (Figure 2.3.2a). Faults are often filled with highly weathered
materials, varying from extremely soft clay and highly fractured and crushed rocks.

Figure 2.3.2a A dyke intrusion.

As discussed earlier, large geological discontinuity features, such as fault zones and
intrusions, are dealt separately. Therefore rock discontinuities in considered in a rock
mass are discontinuities of scale comparable to the rock mass, i.e., joints and fractures.

Rock discontinuities are often associated with groundwater flow, as they acts as flow
channels for groundwater. Groundwater has significant influence on the properties of the
rock mass and hence is should be considered in the rock mass characteristics.

2.3.3 Role of Joints in Rock Mass Behaviour

Rock joints change the properties and behaviour of rock mass in the following terms:
(i) Cuts rock into slabs, blocks and wedges, to be free to fall and move (Figure 2.3.3a);
(ii) Acts as weak planes for sliding and moving;
(iii) Provides water flow channel and creates flow networks (Figure 2.3.3b);
(iv) Gives large deformation;
(v) Alters stress distribution and orientation;

Because the rock materials between rock joints are intact and solid, they have relative
small deformation and low permeability. It is therefore obvious that rock mass behaviour
by large is governed by rock joints.

Figure 2.3.3a Block and wedge sliding and falling due to jointing.

Figure 2.3.3b Groundwater flow out of a joint in a tunnel.


Chapter 2 Rock Formation and Rock Mass 12

2.4 Inhomogeneity and Anisotropy

2.4.1 Inhomogeneity of Rock Materials

Inhomogeneity represents property varying with locations. Most of the engineering


materials have varying degrees of inhomogeneity.

Rocks are formed by nature and exhibits great inhomogeneity.

Figure 2.4.1a shows the texture of some rock materials. Most of the rocks are not
homogeneous in their physical appearance, due to:
(i) different minerals in a rock,
(ii) different bounding between minerals,
(iii) existence of pores,
(iv) existence of microcracks.

Figure 2.4.1a Textures of some common rocks, (a) granite, (b) sandstone, and (c) gneiss.

Inhomogeneity is the cause of fracture initiation leading to the failure of a rock material
(Figure 2.4.1b). A high degree inhomogeneity means some elements in the rock material
matrix are very weak, and therefore usually lead to low overall strength of the rock
material. For two rock specimens of same composition, one contains a pore or a default
usually fails at lower strength than the other.

Figure 2.4.1b Effects of inhomogeneity on failure of rock material.

2.4.2 Inhomogeneity of Rock Masses

Inhomogeneity of a rock mass is primarily due to the existence of discontinuities.

Rock masses are also inhomogeneous due to the mix of rock types, interbedding and
intrusion (Figure 2.4.2a).

Figure 2.4.2a Sedimentary formation with bedding planes and layers.

2.4.3 Anisotropy

Anisotropy is defined as properties are different in different direction.

Anisotropy occurs in both rock materials and rock mass.


Chapter 2 Rock Formation and Rock Mass 13

Some sedimentary rocks, e.g., shale, have noticeable anisotropic characteristics. Other
sedimentary may not have clear anisotropy. However, under the influence of formation
process and pressure, small degree of anisotropy is possible.

Rock with most obvious anisotropy is slate. Phyllite and schist are the other foliated
metamorphic rocks that exhibit anisotropy, as seen in Figure 2.4.3a.

Figure 2.4.3a Some common anisotropic rocks, (i) slate and (ii) shale.

Rock mass anisotropy is controlled by (i) joint set (Figure 2.4.3b), and (ii) sedimentary
layer (Figure 2.4.2a).

Figure 2.4.3b A granitic rock mass with one dominating joint set.

2.5 In Situ Rock Stresses

2.5.1 Overburden Stress and Tectonic Stress

Rock at depth is subjected to natural stress generated by the overlaying rocks and tectonic
force.

Consider a point at a depth (z) below the ground, vertical stress at that point σv is given
by weight of the overlying material, i.e., σv = γ z, where γ is the unit weight of the
overlaying material.

The average specific gravity of rocks is 2.7. The average vertical stress at depth can be
estimated as

σv ≈ 0.027 z

where σv is in MPa, when z is in m.

It is often considered that the horizontal stresses are induced by the vertical stress. But
this suggestion is not correct, as tectonic activities contribute greatly to the built up of
horizontal stresses, perhaps simultaneously. Horizontal stresses in rock are primarily
tectonic.

This is evidently by the fact that horizontal stresses in rocks are generally higher than
vertical stresses. The maximum horizontal stress is the same directions as tectonic
movement, as shown in the stress maps of Europe (Figure 2.5.1a)and of the World.
Tectonic stress has huge variations in term of magnitude. Exceptional high horizontal
stress has been reported in areas of tectonic movement and close to tectonic boundaries.
Chapter 2 Rock Formation and Rock Mass 14

Figure 2.5.1a In situ stress map of Europe.

In situ stress field can also be altered by geological factors and processes:
Surface topography
Erosion
Intrusion
Fault and faulting

2.5.2 In situ Stress Measurements

of vertical stress with depth is scattered about the tend line, σv = 0.027 z, which represents
In situ stress measurements have been compiled and presented in Figure 2.5.2a. Change

the overburden pressure.

Figure 2.5.2a In situ stress measurements at various (Brady Brown 157).

The horizontal stresses are presented in the figure by a ratio of average horizontal stress to

3000 m. (100/z + 0.3) ≤ k ≤ (1500/z + 0.5) is suggested as the limits of k. It should be


vertical stress, k. From the data, k various between 0.5 to more than 3.0, at depth up to

noted, at common depth for civil engineering works, say, less than 1000 m, the variation
of horizontal stress is wide. The maximum horizontal stress could be up to 10 times the
vertical stress.

It is very common in rock mechanics that one of the horizontal stresses represent the
major principal stress, while the vertical stress or the other horizontal stress represents the
minor principal stress.

While vertical stress can be estimated with reasonable reliability. The horizontal stress
should not be estimated. For projects that maximum stress direction and magnitude may
be important, in situ stress measurements is required.

2.5.3 Effective Stress

The principle of effective stress used in soil mechanics has limited application in rock
mechanics. In porous rocks, e.g., sandstone, porewater pressure may be evenly
distributed and hence the effective stress may be computed as total stress – pore pressure.
However, in fractured rock mass, distribution of water is no longer even and stress field is
no longer uniform. Hence, the effective stress principle has no application, as illustrated
in Figure 2.5.3a.
Chapter 2 Rock Formation and Rock Mass 15

Figure 2.5.3a Pore water and effect stress in porous rock and fractured rock mass.

2.5.4 Redistribution of Stress

One major characteristics of rock engineering is that the engineering activities, e.g.,
cutting out a slope or a tunnel, disturbs the original stress field which is already in
equilibrium (for millions of years). Rock mechanics study therefore has to constantly
deal with stress re-distribution and redistributed stresses, as well as the short term
response of rock during stress re-distribution and long term behaviour in the redistributed
stress field. Figure 2.5.4a gives an illustration on how excavating a tunnel causes stress to
redistribute, while Figure 2.5.4b shows the displacement of a rock slope due to stress
release.

Figure 2.5.4a Redistribution of stress due to excavation.

Figure 2.5.4b Movement of rock slope due to stress release at slope face.

2.6 Groundwater

2.6.1 Water in Rock Material

Most of the igneous and metamorphic rocks are very dense with interlocked texture. The
rocks therefore have extremely low permeability and porosity. They are generally
considered as impermeable. Some clastic sedimentary rocks, typically sandstones, can be
porous and permeable.

2.6.2 Flow in Rock Fracture Network

While rock materials may be extremely low in porosity and permeability, the rock masses,
on the other hand, are often fractured. Fractures are opening in the rock masses therefore
provide flow paths. Flow quantity and hydraulic conductivity of fractures can be
estimated by using appropriate hydraulic equations, e.g., Darcy’s Law.

Flow in a fractured rock mass is controlled by the connectivity of fracture system or


network. As shown in Figure 2.6.2a, only a limited percentage of fractured are
interconnected, flow therefore can only be found in those connected fractures while those
non-connected fractures do not have flow. It is commonly observed at site that only a few
fractured has water flow while the remaining fractures are dry.

Figure 2.6.2a Flow in a fractured rock mass, simulated by computer.


Chapter 2 Rock Formation and Rock Mass 16

2.6.3 Effects of Groundwater and Pressure

Groundwater represents an important boundary condition of the rock mass. It is no only


the water pressure which contributes to the stress field. The presence of water, changes a
number of rock parameters, e.g., a wet surface has low friction. The presence of water,
also increase the complexity of rock engineering, e.g., it is more difficult to tunnel with
water inflow and high water pressure.

In the later chapters, the effects of groundwater on the rock mass properties and rock mass
classifications will be addressed again.

2.7 Special Rocks

2.7.1 Weathering and Weathered Rocks

All rocks disintegrate slowly as a result of mechanical weathering and chemical


weathering. Mechanical weathering is the breakdown of rock into particles without
producing changes in the chemical composition of the minerals in the rock. For example,
ice is an important agent of mechanical weathering. Water percolates into cracks and
fissures within the rock, freezes, and expands. The force exerted by the expansion is
sufficient to widen cracks and break off pieces of rock. Heating and cooling of the rock,
and the resulting expansion and contraction, also aids the process. Mechanical weathering
contributes further to the breakdown of rock by increasing the surface area exposed to
chemical agents.

Chemical weathering is the breakdown of rock by chemical reaction. In this process the
minerals within the rock are changed into particles that can be easily carried away. Air
and water are both involved in many complex chemical reactions. The minerals in
igneous rocks may be unstable under normal atmospheric conditions, those formed at
higher temperatures being more readily attacked than those which formed at lower
temperatures. Igneous rocks are commonly attacked by water, particularly acid or alkaline
solutions, and all of the common igneous rock forming minerals (with the exception of
quartz which is very resistant) are changed in this way into clay minerals and chemicals in
solution. Figure 2.7.1a gives the condition of weathering and Figure 2.7.1b shows a
weathered granite.

Figure 2.7.1a Relationship between climate and type of weathering.

Figure 2.7.1b Weathering of granite in a outcrop and weathered granite.


Chapter 2 Rock Formation and Rock Mass 17

Weathering is progressive, between fresh rock and completed weathered and residual
materials (soil), rocks can be slightly, moderately and higly weathered. Those weathered
rocks are still intact and have structure and texture as rock. However, due to weathering,
their properties have been affected and altered. Table 2.7.1a gives the classification of
weathering grade and common properties.

Table 2.7.1a Classification of weathering grade and common properties.


Index Properties
Weathering Description of Rock Material and
Class/Grade Rock Mass RQD Rock/soil Effective
Strength
(%) (% rock) porosity

No visible sign of material weathering.


Fresh Near boundary with Grade II some 90 ~ 95 ~
Very high
(I) slight discoloration on major defects 100 100

Discoloration indicates weathering of Very high to 5%


Slightly
rock material and defect surface. 50~60% of 75 ~ 90 ~ increase
weathered
Discoloration ranges from defect fresh rock 90 95 from fresh
(II)
surface only to completely stained strength rock
Less than 50% of material
Moderately
decomposed and disintegrated to 30% of fresh 40 ~ 60 ~ 7%
weathered
intact soil. Rock core discoloured and rock strength 75 90 increase
(III)
weakened
More than 50% of material
Highly
decomposed and disintegrated to 15% of fresh 10 ~ 30 ~ 10%
weathered
intact soil. Rock core discoloured and rock strength 40 60 increase
(VI)
weakened
Completely Intact friable soil which may be weakly
0~ 0~ 20%
weathered cohesive. Soil has fabric of parent Extremely low
10 30 increase
(V) rock
Residual soil Friable soil with original rock fabric
Extremely low 0 0 > 20
(VI) completely destroyed

The strength of rock materials is governed by the bonding strength of mineral grains in
the rock. Once being weathered, particularly by chemical weathering, the bonding is
weakened and resulting in significant reduction of rock material strength.

2.7.2 Soft Rocks and Hard Soils

Sedimentary rocks are formed by sediments (soils) through lithification (solidification


through compact and cementation). The lithification process is long, and due to
geological movement, the formation could be pushed upward to the surface and
lithification could stoped before the sediments are being completed solidified. The
materials then could be highly consolidated but not fully solidified. Typically, those
materials have low strength and high deformability, and when placed in water, they often
can be dissolved. Figure 2.7.2a shows a semi-lithified sandstone with upper portion being
weathered. In dry state, the material behaves as weak rock and when place in water, it
collapses.
Chapter 2 Rock Formation and Rock Mass 18

Figure 2.7.2a A semi-lithified alluvium sandstone with weathered upper portion.

Such materials some exhibit the properties of rock and of soil. Depending on the
solidification degree, the materials can be described and dealt accordingly, with soil
mechanics principles or rock mechanics principles, or a combination of both.

2.7.3 Swelling Rocks

Some rocks have the characteristics of swelling, that is when the rock is exposed with
water (directly in contact with water or in air), it expanse. This is primarily due the
swelling behaviour of the minerals of the rock, typically the montmorillonite clay mineral.
This montmorillonite mineral is composed of units made of two silica tetrahedral sheets
with a central alumina octahedral sheet. In stacking of these combined units one above the
other, oxygen layers of each unit are adjacent to oxygen of the neighbouring units with a
consequence that there is a very weak bond and an excellent cleavage between them.
Water can enter between the sheets, causing them to expand significantly. Thus rock and
soils containing considerable amount of montmorillonite minerals will exhibit high
swelling and shrinkage characteristics.

Swelling can be characterised by swelling measurement. In engineering, swelling is a


complex problem due to excess deformation when the rock is exposed (Figure 2.7.3a).

Figure 2.7.3a Excessive deformation observed at a tunnelling site sue to swelling.

2.7.4 Highly Fractured and Crushed Rocks

Due to various geological processed, rocks can be highly fractured and sometimes
crushed. Highly fractured and crushed rock masses can exist in large extend due to
regional tectonic stress and movement, or localised in fault and shear zones. Such rock
can often have a mixed sizes and can be a rock-soil mixture. Figure 2.7.4a shows some
typical highly fractured and crushed rock.

Figure 2.7.4a Some typical highly fractured and crushed rock.

Characteristics of such highly fractured and crushed rocks are quite different from the
massive rock mass. They behave as granular and block material. Their mechanical
properties are depending on the geometry and friction. When such materials are
encountered in engineering, they need to be addresses separately.
Chapter 2 Rock Formation and Rock Mass 19

Further Readings

Blyth, FGH, de Freitas MH (1984). A geology for engineers, 7th Edition. Arnold,
London.

Parriaux A (2006). Géologie: bases pour l’ingénieur. PPUR, Lausanne.


Magma cr
ys
(Lava) t al
lis

g
ti n
a tio

el
n

m
Metamorphic metamorphism Igneous
Rock Rock

we
phism

weath
ath

transp
eri
ng
/t ran
mor

ering
ort
sp
ort
meta

Sedimentary weathering/transport Sediment


Rock (Soil)

lithification
2.1.1a

2.1.2a, b

1
2.1.3a

Sandstone

Rock salt

Shale

Conglomerate

2.1.4a

2
Quartzite
Slate Schist Gneiss

2.1.5a

2.1.6a

3
2.1.6b

2.2.1a

4
2.2.1b

2.2.2a

5
2.2.3a

2.2.4a

6
A borehole 10 cm.

A tunnel of 12 m diameter.

An excavated quarry slope of several 10 m high.

2.3.1a

2.3.1b

7
2.3.2a

2.3.3a

8
2.3.3b

Granite Gneiss

Sandstone

2.4.1a

9
2.4.1b

2.4.2a

10
2.4.3a

2.4.3b

11
2.5.1a

2.5.2a

12
Pore
water

2.5.3a

σV σV
σV σV

2.5.4a

13
2.5.4b

2.6.2a

14
2.7.1a

Weathered

Fresh

Fresh

Weathered

2.7.1b

15
2.7.2a

2.7.3a

16
2.7.4a

17
Chapter 3 Principe of Mechanics 1

CHAPTER 3
PRINCIPLES OF MECHANICS
This Chapter is intended as a review on the mechanics principles that relevant to rock
mechanics. Reader who is not familiar with mechanics should start with one of the
many textbook on mechanics, listed at the end of this Chapter.

3.1 Stress, Strain and Deformation Modulus

3.1.1 Normal and Shear Stresses

Stress is commonly defined the internal distribution of force per unit area that balances
and reacts to external loads applied to a body. It is a second-order tensor with nine
components, but can be fully described with six components due to symmetry in the
absence of body moments (Wikipedia).

Stresses can be divided down into normal stresses and shear stresses. Normal stress is
defined as stress perpendicular to the plane where the stress is act on. Shear stress is the
stress parallel to the plane where the stress is act on. They are perhaps best represented
by the diagram shown in Figure 3.1.1a.

Figure 3.1.1a Normal and shear stresses on an infinitesimal cube aligned with the
Cartesian axes. (Hudson 35)

The stresses act on an infinitesimal cube aligned with the Cartesian axes has nine

σxx, σyy and σzz; and the six shear stresses are τxy, τxz, τyx, τyz, τzx, and τzy.
components: three normal stresses and six shear stresses. The three normal stresses are

Often, the nine stress components are often conveniently represented in matrix form,
where the row representing the stresses on a plane and the column representing the
stresses in a direction,

σxx τxy τxz


σ = τyx σyy τyz
τzx τzy σzz
(3.1.1a)

By considering the state of equilibrium, each pair of shear stresses are equal, hence,

τxy = τyx, τxz = τzx, τyz = τzy (3.1.1b)

The stress components in the matrix can be reduced from nine to six, and the stress matrix
can be written as,
Chapter 3 Principe of Mechanics 2

σxx τxy τxz


σ = τxy σyy τyz
τxz τyz σzz
(3.1.1c)

3.1.2 Principal Stresses

Usually, there are nine stress components acting in three directions on three planes.
However, if the planes where the stresses act on are rotated, there is a position that the
magnitude of shear stresses is zero. Hence, the stresses acting on the three planes are
only normal stresses. In this situation, the three normal stresses are termed as principal
stresses.

σ1 0 0
σ = 0 σ2 0
0 0 σ3
(3.1.2a)

Figure 3.1.2a shows the principle of stress transformation on a 2D element subjected to

a-b plane is σab and τab.


normal and shear stresses. Considering a plane a-b, the normal and shear stresses on the

Figure 3.1.2a Stress transformation and Mohr’s circle representation.

Summing the components of forces acting on the element in the direction of a-b plane, it
gives, in the normal direction,

σab (ab) = σx sinθ (ab) sinθ + σy cosθ (ab) cosθ + 2 τxy (ab) sinθ cosθ

or

σab = σx sin θ + σy cos θ + 2 τxy sinθ cosθ


2 2
(3.1.2b)

or

σy + σx σy – σx
σab = + cos2θ + τxy sin2θ (3.1.2c)
2 2

here, σy is assumed to be greater than σx.

Again, in the shear direction along a-b plane, the force equilibrium can be written as,

τab (ab) = -σx cosθ (ab) sinθ + σy sinθ (ab) cosθ – τxy cosθ (ab) cosθ + τxy sinθ (ab) sinθ
Chapter 3 Principe of Mechanics 3

or

τab = σy sinθ cosθ – σx sinθ cosθ – τxy (cos θ – sin θ)


2 2

or

σy – σx
τab = sin2θ – τxy cos2θ
2

To make the a-b plane that only principal stress exist, i.e., τab = 0, then,

2 τxy
σy – σx
tan2θ =

The above equation gives two values of θ that are 90° apart. This means that there are in
fact two planes at right angles to each other on which shear stress is zero. These planes
are the principal planes and stresses acting on these planes are principal stresses. By
substituting the above equations, the values of the principal stresses can be found,

σy + σx σ – σx 2
σab = σ1 = ) + τxy ]
2 ½
+ [( y
2 2

as the major principal stress, and,

σy + σx σ – σx 2
σab = σ2 = ) + τxy ]
2 ½
– [( y
2 2

as the minor principal stress.

The normal stress and shear stress acting on any plane can also be determined by using
the Mohr’s circle, as shown in Figure 3.1.2a. On the Mohr’s circle, the points where the
circle intercept the normal stress axis, where shear stress is zero, are the values of
principal stresses.

The Mohr’s circle can also be used to determine normal and shear stresses acting on a
plane of a given angle to the given principal stresses value and direction. This will be
illustrated later in dealing with the Mohr-Coulomb strength criterion.

The concept of principal stresses is important in rock mechanics. For example, in


excavation, the excavated face is free from shear stress and normal stress, therefore, the
rock at this position is subjected to principal stresses and one (normal to the excavated
face) of the principal stresses is in fact zero.

3.1.3 Deformation and Strain


Chapter 3 Principe of Mechanics 4

When a material is subject to stress, the material deforms. The deformation occurs not

Again, refer the cube in to Figure 3.1.3a, with the normal stress, σxx, the cube deforms
only in the direction of the stress, but also in the other two perpendicular directions.

(contracts) in the direction of x-x, and expends in the directions of y-y and z-z.

Figure 3.1.3a Deformations due to stress on an infinitesimal cube.

Strain is defined as the ratio of deformation to the original dimension, i.e.,

ε=δ/l (3.1.3a)

It therefore has no dimension. Often, it is expressed in term of percentage.


Strain can also expressed in matrix form:

εxx γxy γxz


ε = γxy εyy γyz
γxz γyz εzz
(3.1.3b)

εxx, εyy and εzz are normal strain, and γxy, γxz and γyz are shear strain.

Strain is a mechanical property of the material, it only occurs when a stress is applied,
directly or through other means. In another word, when there is a strain, there must be
stress accompanying the strain.

Often, volumetric strain is used in rock mechanics. It is defined as the ratio of volume
change to the original volume, i.e.,

εv = Δv / v (3.1.3c)

3.1.4 Poisson’s Ratio

When a sample of material is stretched in one direction, it tends to get thinner in the other
two directions. Poisson's ratio is a measure of this tendency. Poisson's ratio the ratio of the
contraction strain (normal to the applied load) divided by the extension strain (in the
direction of the applied load). Refer to Figure 3.1.3, with a given normal stress in x-x
direction, strains occur in the direction of x-x, as well as in the directions of y-y- and z-z.
The ratio of strain in y-y- or z-z direction to strain in x-x direction is defined as the
Poisson’s ratio,

ν = εyy / εxx, ν = εzz / εxx (3.1.4a)

For a perfectly incompressible material, the Poisson's ratio would be exactly 0.5. Most
rock materials have ν between 0.2 and 0.4. The Poisson’s ratio is a mechanical property
Chapter 3 Principe of Mechanics 5

of the material. For a homogeneous material, the Poisson’s ratios in y-y and z-z
directions are equal.

3.1.5 Deformation Modulus

Deformation modulus is a mechanical property indicating the rate of change of strain with
change of stress. For an elastic material, it is termed as Modulus of Elasticity, and often
referred as the Young’s Modulus. It is defined as the ratio, for small strains, of the rate
of change of stress with strain. This can be experimentally determined from the gradient
of a stress-strain curve obtained from loading tests conducted on a sample of the material.

E = Δσ / Δε (3.1.5a)

The Young’s Modulus, E, has the same unit as the stress, σ.

When a material is subject to a stress, it undergoes strain. The stress-strain relationship


can often to be represented by the stress-strain curve, typically shown in Figure 3.1.5a.
The deformation modulus at a specific stress level is the gradient of the stress-strain curve
at point of that specific stress level. For an elastic material, the stress-strain is a straight
line, and the modulus is therefore the gradient of the line.

Figure 3.1.5a Determination of deformation modulus from stress-strain curves.

Stress-strain relationship can be expressed in matrix,

εxx σxx
εyy σyy
S11 S12 S13 S14 S15 S16

εzz σzz
S21 S22 S23 S24 S25 S26

γxy σxy
S31 S32 S33 S34 S35 S36
= (3.1.5b)

γyz σyz
S41 S42 S43 S44 S45 S46

γzx σzx
S51 S52 S53 S54 S55 S56
S61 S62 S63 S64 S65 S66

ε = S σ (3.1.5c)

σ = D ε (3.1.5d)

εxx –ν –ν σxx
εyy –ν 1 –ν σyy
1 0 0 0

εzz –ν –ν 1 σzz
0 0 0

γxy 0 0 2(1+ν) 0 σxy


1 0 0 0
= (3.1.5e)

γyz 0 2(1+ν) 0 σyz


E 0 0

γzx 0 2(1+ν) σzx


0 0 0
0 0 0 0
Chapter 3 Principe of Mechanics 6

3.1.6 Plane Stress and Plane Strain

For a thin wall plate, a 3D problem can be treated as a 2D plane stress problem. The
stresses and strains are expressed as

σxx τxy
σ = τxy σyy
0
0 (3.1.6a)
0 0 0

εxx γxy
ε = γxy εyy
0

εzz
0 (3.1.6b)
0 0

εzz can often temporarily removed from analysis to make all stress and strain only the in-
plane terms, and effective becomes a 2D analysis.

For a long tunnel, a 3D problem can be approximated by a 2D plane strain problem. The
strains and stresses are expressed as:

εxx γxy
ε = γxy εyy
0
0 (3.1.6c)
0 0 0

σxx τxy
σ = τxy σyy
0

σzz
0 (3.1.6d)
0 0

σzz is needed to maintain εzz being zero. σzz stress term can be temporarily removed from
the analysis to leave only the in-plane terms, effectively reducing the 3D problem to a
much simpler 2D problem.

3.2 Strength and failure criteria

3.2.1 Basic Definitions

When a material is subjected to a stress, the material deforms. When the stress increases
to a certain state, the material starts to yield and subsequently will either loss the strength
substantially or will undergo large strain without additional stress, as shown typically in
Figure 3.2.1a.
Chapter 3 Principe of Mechanics 7

Figure 3.2.1a Stress-strain curve with yield point, peak strength, post-peak ductile and
brittle behaviour.

Strength of a material is the limit states of stress (peak or yield) the material can sustain. It
is often considered in terms of compressive strength, tensile strength, and shear strength,
which are the limit states of compressive stress, tensile stress and shear stress
respectively.

There are various definitions of strength, some represented by yield and some by peak.
In rock mechanics, common definition is by peak strength.

Strain occurred at the peak strength is often defined as strain at failure. This is a useful
measure in rock mechanics and is associated with the brittleness of the material.

Hardness is another commonly used term describing rock material. It is the


characteristic of a solid material expressing its resistance to permanent deformation.

Toughness is the resistance to fracture of a material when stressed. It is defined as the


amount of energy that a material can absorb before rupturing, and can be found by finding
the area (i.e., by taking the integral) underneath the stress-strain curve.

3.2.2 Post-Peak Stress-Strain Behaviour

Post-peak behaviour is the stress-strain behaviour after the peak strength. When a
material has reached the limit of its strength, it usually has the option of either fracture or
deformation, approximately resulting to two typical post-peak behaviours: brittle and
ductile, as shown in Figure 3.2.1a.

Brittle behaviour indicates a substantial reduction of stress-carrying capacity, reflecting


fracturing of the material under stress. At post-peak region, the material may have
residual strength but that residual strength is generally substantially lower than the peak
strength. A brittle material usually has undergoes small strain before failure, i.e., small
strain at failure. A brittle failure is generally associated with fracture by tension rather
than shear, and there is little or no evidence of plastic deformation before failure.

Ductile behaviour is represented by being capable of sustaining large deformations with


losing stress carrying capability in the post-peak region. Ductility is the physical
property of being capable of sustaining large plastic deformations without fracture.

Post-peak behaviour can be influenced by pressure and temperature. This happens as an


example in the brittle-ductile transition zone at an approximate depth of 10 km in the
Earth's crust, at which rock becomes less likely to fracture, and more likely to deform
ductilely.
Chapter 3 Principe of Mechanics 8

Residual strength is the stress carrying capacity at post-peak region. This strength
indicates that even after failure, rock often has a (much smaller) load carrying capacity,
which can have significant effects on rock engineering.

3.2.3 Yield Strength and Criteria

Yield strength, or the yield point, is defined in engineering and materials science as the
stress at which a material begins to plastically deform. Prior to the yield point the material
will deform elastically and will return to its original shape when the applied stress is
removed. Once the yield point is passed some fraction of the deformation will be
permanent and non-reversible. Knowledge of the yield point is vital when designing a
component since it generally represents an upper limit to the load that can be applied
(Wikipedia).

A yield criterion is a hypothesis concerning the limit of elasticity under any combination
of stresses. There are two interpretations of yield criterion: one is purely mathematical in
taking a statistical approach while other models attempt to provide a justification based on
established physical principles. Since stress and strain are tensor qualities they can be
described on the basis of three principal directions.

The most common yield criteria are briefly outlines below. Details of those criteria are
readily available in textbooks of mechanics and strength.

(a) Tresca-Guest criterion

The Tresca-Guest criterion is the most simple yield surface. In principal stresses it is
expressed as follows:

max(|σ1 – σ2|, |σ2 – σ3|, |σ3 – σ1|) = σ0

The Tresca-Guest criterion in three dimensional space of principal stresses is a prism of


infinite length and six sides, as illustrated in Figure 3.2.3a. This means that material
remains elastic when all three principal stresses are roughly equivalent (a hydrostatic
pressure), no matter how much compressed or stretched. But when the material is
subject to shearing, one of principal stresses becomes smaller (or bigger), then the yield
surface is crossed and material enters plastic domain. In two dimensional space, it is a
cross diagonally cut section of the prism (Figure 3.2.3b).

Figure 3.2.3a Tresca-Guest criterion in 3D spaces of principal stresses.

Figure 3.2.3b Tresca-Guest and Mises criteria in 2D spaces of principal stresses.

(b) Mises criterion


Chapter 3 Principe of Mechanics 9

The von Mises criterion is expressed as follows:

Also it can be expressed in non-principal stresses as below:

The von Mises criterion in three dimensional space of principal stresses is a circular
cylinder of infinite length, with the same angle to all three axes, as seen in Figure 3.2.3c.
In two dimensional space of principal stresses, it is an ellipse that diagonally cut cross the
cylinder (Figure 3.2.3b).

Figure 3.2.3c Mises criterion in 3D spaces of principal stresses.

(c) Mohr-Coulomb criterion

The Mohr-Coulomb criterion is a first two-parametric yield surface, for the maximum
compression and tension. The model is the first one that takes shearing into account. It is
expressed as follows:

or

The parameters are Rc and Rr which are the maximum values for compression and
tension for the given material. it should be noted that the criterion considers the
maximum difference between the major and the minor principal stresses only, and does
not take the intermediate principal stress in the strength criterion.
Chapter 3 Principe of Mechanics 10

The Mohr-Coulomb criterion in three dimensional space of principal stresses is


represented by a conical prism (Figure 3.2.3d). If K = 0 then it becomes Tresca-Guest
criterion, thus K determines the inclination angle of conical surface. In two dimensional
space of principal stresses, it is a cross section of the conical prism (Figure 3.2.3e).

Figure 3.2.3d Mohr-Coulomb criterion in 3D space of principal stresses.

Figure 3.2.3e Mohr-Coulomb and Drucker-Prager criteria in 2D space of principal


stresses.

The Mohr-Coulomb strength criterion can be represented graphically, by Mohr’s circle.

Most of the classical engineering materials, including rock materials, somehow follow
this rule in at least a portion of their shear failure envelope. Application and discussion
of the Mohr-Coulomb strength criterion in rock mechanics is given in Chapter 4.

(d) Drucker-Prager criterion

The Drucker-Prager criterion is expressed as follows:

cone (Figure 3.2.3f). If α = 0 then it becomes the Mises criterion. In two dimensional
The Drucker-Prager criterion in three dimensional space of principal stresses is a regular

space of principal stresses, it is a cross section of this cone, which produces an ellipsioidal
shape (Figure 3.2.3e).

Figure 3.2.3f Drucker-Prager criterion in 3D spaces of principal stresses.

(e) Unified Strength Criterion


Chapter 3 Principe of Mechanics 11

The Unified Strength Criterion by Yu has the following characteristics:


(i) It is able to reflect the fundamental characteristics of brittle materials (including rock
and concrete), i.e., different tensile and compressive strengths, hydrostatic pressure
effect, the effect of intermediate principal stress and its zonal change and material
dependence.
(ii) It has a clear physics and mechanics background, a unified mathematical model, and a
simple and explicit criterion, which includes all independent stress components and
simple material parameters.
(iii) It is also suitable for different types of brittle materials under various stress states, and
is consistent with the triaxial test results.
(iv) It can be easily applied to analytical and numerical modeling.

The Unified Strength Criterion expressed in terms of three principal stresses is as follows:

σ1 – α (b σ2 + σ3)/(1 + b) = σt, when σ2 ≤ (σ1 + α σ3)/(1 + α)

(σ1 + b σ2)/(1 + b) – α σ3 = σt, when σ2 ≥ (σ1 + α σ3)/(1 + α)

where σt is tensile strength, α is the strength ratio of tensile to compressive strength


(σt/σc), b is the intermediate principal stress parameter.

When b = 0, the Unified Strength Criterion becomes the Mohr-Coulomb (σ1 – α σ3 = σt).

of b varies between 0 and 1 (0 ≤ b ≤ 1), to reflect the characteristics of various different


The Unified Strength Criterion can produce a full spectrum of new criteria when the value

materials. The Unified Strength Criterion is especially versatile in reflecting the σ2


effect to different extents for different materials.

In general, the failure surface of the Unified Strength Criterion is a dodecahedral-shaped


cone about the hydrostatic axis. The failure surface in the plane perpendicular to the

0 ≤ b ≤ 1. When b = 0 or 1, the dodecahedral-shaped cone reduces to hexagonal. When


cone axis is shown in Figure 3.2.3g. It is quite obvious that the surface is convex when

α=1, the Unified Strength Criterion is simplified to the Unified Yield Criterion applicable
to these materials with the same yield stress in tension and compression, and the three-
fold symmetric limit surfaces are simplified to six-fold symmetric yield surfaces.

Figure 3.2.3g Failure surfaces of the Unified Strength Criterion on deviatoric plane.

3.3 Fracture Mechanics

3.3.1 Fracture Initiation and Propagation


Chapter 3 Principe of Mechanics 12

Fracture mechanics is a sub-division of solid mechanics studying the failure of a material


containing a crack, in order to understand the initiation and propagation of cracking. It
uses methods of analytical solid mechanics to calculate the driving force on a crack and
methods of experimental solid mechanics to characterize the material's resistance to
fracture.

Fracturing of a material has to start somewhere within the material. It usually starts at a
location of stress concentration. In most materials, they cannot be perfectly
homogeneous and there are often existing pores and cracks within the material formed at
its creation or through various processes. For example, all rock materials contain micro-
cracks and pores. When loaded, those existing cracks and pores causes stress
concentration at their tips and from there, new cracks will be initiated and further
propagated.

There are three types of crack initiation and propagation (Figure 3.3.1a):
Mode I crack – Opening mode (a tensile stress normal to the plane of the crack), this
mode is most relevant to rock mechanics.
Mode II crack - Sliding mode (a shear stress acting parallel to the plane of the crack and
perpendicular to the crack front);
Mode III crack - Tearing mode (a shear stress acting parallel to the plane of the crack and
parallel to the crack front).

Figure 3.3.1a Three modes of crack initiation and propagation.

3.3.2 Griffith Crack Theory

The Griffith crack theory is a theory of brittle fracture, using elastic strain energy
concepts. It describes the behaviour of crack propagation of an elliptical nature by
considering the energy involved. The equation basically states that when a crack is able to
propagate enough to fracture a material, that the gain in the surface energy is equal to the
loss of strain energy, and is considered to be the primary equation to describe brittle
fracture. Because the strain energy released is directly proportional to the square of the
crack length, it is only when the crack is relatively short that its energy requirement for
propagation exceeds the strain energy available to it. Beyond the critical Griffith crack
length, the crack becomes dangerous.

For the simple case of a thin rectangular plate with a crack perpendicular to the load
Griffith’s theory becomes:

G=πσ a/E
2

G c = π σf a / E
2

where G is the strain energy release rate, σ is the applied stress, a is half the crack length,
and E is the Young’s modulus. The strain energy release rate can otherwise be understood
Chapter 3 Principe of Mechanics 13

as the rate at which energy is absorbed by growth of the crack. When G is greater than
the critical value Gc, crack will begin to propagate.

This concept is often referred as Griffith’s energy instability concept. This concept was
applied in rock mechanics to develop the Griffith strength criterion, as outlined in te next
chapter.

3.3.3 Stress Intensity and Fracture Toughness

A modification of Griffith’s theory was made by Irwin. In the modified criterion, strain
energy release rate is replaced by stress intensity (KI)and surface energy is replaced by
fracture toughness (Kc). Both of these terms are simply related to the energy terms in the
original Griffith theory:

KI = σ (π a)½

Kc = (E Gc) ½ for plane stress

Kc = [E Gc /(1 – ν)]½ for plane strain

where ν is the Poisson’s ratio. It is important to note that Kc has different values when
measured under plane stress and plane strain conditions

Fracture occurs when KI ≥ Kc. For the special case of plane strain deformation, Kc
becomes KIc and is considered a material property. The subscript I refers to mode I
cracking.

3.4 Mechanics of Discontinuity

3.4.1 Discontinuity and Global Continuum Law

Discontinuity in a continuous material is defined a plane feature where continuity of the


materials is no longer hold. They are usually significant mechanical breaks in a material.
Continuity may continue to exist after the discontinuity. Typical discontinuities are
fractures, interfaces and joints.

In rock, discontinuities are fractures, joints, faults and bedding plane. While the rock
materials are often considered as continuous materials, rock fractures and joints are
commonly treated as discontinuities.

While the mechanics at discontinuity is discontinuous, certain laws of continuum are still
hold.

3.4.2 Stress and Strain at Discontinuity


Chapter 3 Principe of Mechanics 14

Stresses are often disturbed by a discontinuity. For example, at non-fully-contacted


fracture, opening exists. Normal stress on the opening is zero and there are stress
concentrations on the contact points of the fracture surface. The stress field is no longer
the same as in the continuous material.

The only exception is a fully-contact plane perpendicular to a uniaxial load. In this case,
the stress field is continuous although strain may not.

Similarly, displacement at discontinuity is not continuous. For example, at a fracture


plane, sliding or shear displacement may occur. There may be much greater normal
displacement at fracture than those of the material.

Discontinuities can range from a fully-contacted interface to an opening containing


different material. The mechanics of are vary different.

For a fully-welded interface between two different materials, the interface actually has the
continuities both is stress and displacement. The discontinuity is presented in the change
of mechanics of the two materials on each side of the interface.

For a fully-contacted smooth interface, the interface is subjected to shear displacement.

For a locally-contacted fracture, i.e., there are void between the two sides, both stress and
displacement discontinuous are expected.

3.4.3 Mechanics of Discontinuities for Normal Stress and Deformation

Normal stress and displacement of fully-contact discontinuity is continuous and therefore


can be dealt with continuum approach.

For a locally-contacted fracture shown in Figure 3.4.3a, i.e., there are void between the
two sides, stress-displacement function is discontinuous. For an idealised pillar-like
locally-contacted fracture shown below, assume the ratio of contact area to intact area (A)
is k (k=0~1), i.e., contact area at fracture is kA.

Figure 3.4.3a Idealised fracture with local contact of constant area.

Assume the normal stress in the intact portion is N and the total load is NA. The normal
stress in the contact area is NA/mA = N/m. Since 0<m≤1, stress in the contact area is
always higher than that in the intact portion. Assume the material modulus (E) is the
same for the intact portion and the contact area, then strain of the contact area is N/kE.
If there is no damage of the contact element, the contact area will remain the same with
change of load. Therefore, the load-deformation of the contact zone is linear. Stiffness
(k) of the contact zone is a constant, is given as,
Chapter 3 Principe of Mechanics 15

Load NA mAE
k= = = (3.4.3a)
Deformation (N / m E) d d

For a prism shape locally-contacted fracture shown in Figure 3.4.3b, the contact area has
an initial contact area of A0, and a thickness of d. The contact area in this case is no

zone. i.e., ΔA∝Δh2.


longer constant; it in fact increases with increasing displacement (closure) of the contact

Figure 3.4.3b Idealised fracture with local contact of changing area.

With increase of loading, and discontinuity closes, the contact area increases. Therefore,
at beginning of loading, the contact area is small, rate of deformation is large. At higher
load, the contact area increases, the rate of deformation becomes smaller. Clearly the
load-displacement of the contact zone is no longer linear. This non-linear load-
displacement characteristics is typical for rock fractures, as shown in Figure 3.4.3c.

Figure 3.4.3c Typical non-linear load-displacement characteristics of rock fractures.

3.4.4 Mechanics of Discontinuities for Shear Stress and Deformation

The most common known shear phenomenon of a discontinuity is the sliding between

between the friction angle φ, the normal force (N) and shear force (S), as S = N tanφ.
two contact surfaces (Figure 3.4.4a), i.e., the friction theory. It gives the relationship

Figure 3.4.4a Sliding between two horizontal contact surface.

When slipping at the surface of contact is about to occur, the maximum static frictional
force is proportional to the normal force. When slipping is occurring, the kinetic
frictional force is proportional to the normal force, as shown in Figure 3.4.4b.

Figure 3.4.4b Mobilisation of shear stress with applied shear load.

If the contact surface is at an inclined up angle (i), shown in Figure 3.4.4c, from the force
diagram, along the sliding direction, the normal force is N cos(i) + S sin(i), the shear force
is S cos(i) – N sin(i).

Figure 3.4.4c Sliding between two contact surface at an inclination.


Chapter 3 Principe of Mechanics 16

By friction theory,

S cos(i) – N sin(i) = [N cos(i) + S sin(i)] tanφ,

S – N tan(i) = N tanφ + S tanφ tan(i),

S = N [tanφ + tan(i)] / [1 + tanφ tan(i)],

S = N tan(φ+i)

3.5 Flow Mechanics of Porous Material and Parallel Plates

3.5.1 Flow in Porous Materials

Consider a cylindrical sample of porous material (e.g., soil or rock) under the different
water pressure heads h1 and h2, shown in Figure 3.5.1a. In one dimension, steady water
flows through the fully saturated sample without affecting the structure of the soil or rock,
in accordance with Darcy's flow law,

Q=Aki (3.5.1a)

where Q = volume of water flowing per unit time, A = cross sectional area of sample
corresponding to the flow, k = coefficient of permeability, i = hydraulic gradient = (h1–
h2)/L, and L = length of sample.

Figure 3.5.1a Darcy's flow experiment on porous material.

Hence

Q QL
k= = (3.5.1b)
Ai A (h1 – h2)

The coefficient of permeability k is a constant. Experiments have shown, however, that


its value depends not only upon the character of the material, but also upon the properties

kinematic viscosity, ν, which can be expressed as,


of the fluid percolating through it. The value of k is inversely proportional to fluid

ν
Kg
k= (3.5.1c)

where g = gravitational acceleration, K = the intrinsic permeability, and is a property of


the material only, with dimension of L2.
Chapter 3 Principe of Mechanics 17

3.5.2 Flow between Parallel Plates

For flow of a viscous fluid through a narrow interspace between two closely spaced
parallel plates (e.g., rock fractures), shown in Figure 3.5.2a, the Darcy’s flow law is
applicable when the flow is laminar.

Figure 3.5.2a Darcy's flow experiment on parallel plates.

The intrinsic permeability for laminar flow between parallel plates is (Todd 1959; Verruijt
1970),
2
d
K= (3.5.2a)
12

where d is the thickness of the fluid lamina, i.e., the aperture of the two parallel plates.

Hence for laminar flow through smooth parallel plates, flow equation can be written as
2

12 ν
gd
k= (3.5.2b)

This is often called the "parallel plate theory" in the flow mechanics of rock joints. A

≤2300.
laminar flow through smooth parallel plates is a potential flow with its Reynolds number

By combining the above equation with Darcy’s equation, it gives

12 ν
Aigd
Q= (3.5.2c)

Since the area A is the flow passage which is equal to the width w times to the aperture of
the parallel plates, d, the above equation may be written
3

12 ν
wigd
Q= (3.5.2c)

The above equation is identified as the "cubic flow law" which is widely used to describe
the flow of fluid through parallel plates. It is also used to describe flow in rock joints.

For a planar array of parallel smooth openings, the equivalent permeability parallel to this
array is given as,
Chapter 3 Principe of Mechanics 18

3
d
k= (3.5.2d)
12 b

where b is the spacing between openings.

The above equations show that the flow rate and permeability are extremely sensitive to
the aperture of the opening.

3.6 Empirical Approaches

3.6.1 Use of Empirical Equations

Empirical approaches are often used when theoretical approaches are limited due to many
reasones, for example, the complexity of the material which leads to non-conformable to
theory. Empirical equations then are obtained usually based on extensive experiment
results, by regression. New regression techniques, e.g., neural network, have also been
applied to analysis large amount variables and data.

Empirical criteria commonly used in rock mechanics are the Hoek-Brown criterion for
both rock material and rock mass, JRC-JCS shear strength equation for rock fractures
(Figure 3.6.1a) and several others.

Figure 3.6.1a Empirical JRC-JCS shear strength criterion.

3.6.2 Linear and Multiple Regression

3.6.3 Neural Network and other New Methods

Further Readings

Gere JM, Timoshenko SP, Mechanics of Materials, 2nd Edition. PWS-Kent, Boston,
1984.

Jaeger JC, Cook NGW, Fundamentals of Rock Mechanics, 3rd Edition. Chapman and
Hall, London, 1979.
σz
y x

τxx
z
τzy
τxz
τyz
τxy σx
τyx

σy

3.1.1a

½ (σx + σy) √[½(σx+σy)]2+ τxy2

τ
σy σy, τxy
τxy

σx σab
S σ2 σ1 M σn
b

τxy
O C
τab

θ c

σx, -τxy
a

3.1.2a

1
σx
δy δx

δz

3.1.3a

σ1

ε1
1

3.1.5a

2
σ Peak Strength

Yield Point Ductile

Brittle

3.2.1a

3.2.3a

3
3.2.3b

3.2.3c

4
3.2.3d

3.2.3e

5
3.2.3f

3.2.3g

6
Mode I Mode II Mode III

3.3.1a

Stress N, Area A, Modulus E

Contact area mA, thickness d

3.4.3a

7
Stress N, Area A, Modulus E

Initial contact area Ao,


thickness d

3.4.3b

3.4.3c

8
N S

3.4.4a

F
Fs
Fk
N
P
P
F=

F
N
P

3.4.4b

9
N S
i

3.4.4c

h1
h2
L

Q A

3.5.1a

10
h1
h2
w
L

d
Q

3.5.2a

τ = σn tan[JRC JMC log(JCS/σn+φr)]

3.6.1a

11
Chapter 4 Properties of Rock Materials 1

CHAPTER 4
PROPERTIES OF ROCK MATERIALS
Rock material is the intact rock portion. This Chapter addresses properties of rock
material.

4.1 Physical Properties of Rock Material

4.1.1 Density, Porosity and Water Content

Density is a measure of mass per unit of volume. Density of rock material various, and
often related to the porosity of the rock. It is sometimes defined by unit weight and
specific gravity. Most rocks have density between 2,500nd 2,800 kg/m3.

Porosity describes how densely the material is packed. It is the ratio of the non-solid
volume to the total volume of material. Porosity therefore is a fraction between 0 and 1.
The value is typically ranging from less than 0.01 for solid granite to up to 0.5 for porous
sandstone. It may also be represented in percent terms by multiplying the fraction by
100%.

Water content is a measure indicating the amount of water the rock material contains. It is
simply the ratio of the volume of water to the bulk volume of the rock material.

Density is common physical properties. It is influenced by the specific gravity of the


composition minerals and the compaction of the minerals. However, most rocks are well
compacted and then have specific gravity between 2.5 to 2.8. Density is used to estimate
overburden stress.

Density and porosity often related to the strength of rock material. A low density and high
porosity rock usually has low strength.

Porosity is one of the governing factors for the permeability. Porosity provides the void
for water to flow through in a rock material. High porosity therefore naturally leads to
high permeability.

Table 4.1.1a gives common physical properties, including density and porosity of rock
materials.
Chapter 4 Properties of Rock Materials 2

Table 4.1.1a Physical properties of fresh rock materials.


Rock Dry Density Porosity (%) Schmidt Hardness Cerchar P-Wave Velocity S-Wave Velocity Coefficient of
(g/cm3) Index Abrasivitiy Index (m/s) (m/s) Permeability (m/s)
Igneous
Granite 2.53 – 2.62 1.02 – 2.87 54 – 69 4.5 – 5.3 4500 – 6500 3500 – 3800 10-14 – 10-12
Diorite 2.80 – 3.00 0.10 – 0.50 4.2 – 5.0 4500 – 6700 10-14 – 10-12
Gabbro 2.72 – 3.00 1.00 – 3.57 3.7 – 4.6 4500 – 7000 10-14 – 10-12
Rhyolite 2.40 – 2.60 0.40 – 4.00 10-14 – 10-12
Andesite 2.50 – 2.80 0.20 – 8.00 67 2.7 – 3.8 4500 – 6500 10-14 – 10-12
Basalt 2.21 – 2.77 0.22 – 22.1 61 2.0 – 3.5 5000 – 7000 3660 – 3700 10-14 – 10-12
Sedimentary
Conglomerate 2.47 – 2.76 1.5 – 3.8 10-10 – 10-8
Sandstone 1.91 – 2.58 1.62 – 26.4 10 – 37 1.5 – 4.2 1500 – 4600 10-10 – 10-8
Shale 2.00 – 2.40 20.0 – 50.0 0.6 – 1.8 2000 – 4600
Mudstone 1.82 – 2.72 27 10-11 – 10-9
Dolomite 2.20 – 2.70 0.20 – 4.00 5500 10-12 – 10-11
Limestone 2.67 – 2.72 0.27 – 4.10 35 – 51 1.0 – 2.5 3500 – 6500 10-13 – 10-10
Metamorphic
Gneiss 2.61 – 3.12 0.32 – 1.16 49 3.5 – 5.3 5000 – 7500 10-14 – 10-12
Schist 2.60 – 2.85 10.0 – 30.0 31 2.2 – 4.5 6100 – 6700 3460 – 4000 10-11 – 10-8
Phyllite 2.18 – 3.30
Slate 2.71 – 2.78 1.84 – 3.64 2.3 – 4.2 3500 – 4500 10-14 – 10-12
Marble 2.51 – 2.86 0.65 – 0.81 5000 – 6000 10-14 – 10-11
Quartzite 2.61 – 2.67 0.40 – 0.65 4.3 – 5.9 10-14 – 10-13
Chapter 4 Properties of Rock Materials 3

4.1.2 Hardness

Hardness is the characteristic of a solid material expressing its resistance to permanent


deformation. Hardness of a rock materials depends on several factors, including mineral
composition and density. A typical measure is the Schmidt rebound hardness number.

4.1.3 Abrasivity

Abrasivity measures the abrasiveness of a rock materials against other materials, e.g.,
steel. It is an important measure for estimate wear of rock drilling and boring
equipment.

Abrasivity is highly influenced by the amount of quartz mineral in the rock material.
The higher quartz content gives higher abrasivity.

Abrasivity measures are given by several tests. Cerchar and other abrasivity tests are
described later.

4.1.4 Permeability

Permeability is a measure of the ability of a material to transmit fluids. Most rocks,


including igneous, metamorphic and chemical sedimentary rocks, generally have very low
permeability. As discussed earlier, permeability of rock material is governed by
porosity. Porous rocks such as sandstones usually have high permeability while granites
have low permeability. Permeability of rock materials, except for those porous one, has
limited interests as in the rock mass, flow is concentrated in fractures in the rock mass.
Permeability of rock fractures is discussed later.

4.1.5 Wave Velocity

Measurements of wave are often done by using P wave and sometimes, S waves. P-
wave velocity measures the travel speed of longitudinal (primary) wave in the material,
while S-wave velocity measures the travel speed of shear (secondary) wave in the
material. The velocity measurements provide correlation to physical properties in terms
of compaction degree of the material. A well compacted rock has generally high
velocity as the grains are all in good contact and wave are travelling through the solid.
For a poorly compact rock material, the grains are not in good contact, so the wave will
partially travel through void (air or water) and the velocity will be reduced (P-wave
velocities in air and in water are 340 and 1500 m/s respectively and are much lower than
that in solid). Typical values of P and S wave velocities of some rocks are given in
Table 4.1.1a. Wave velocities are also commonly used to assess the degree of rock mass
fracturing at large scale, using the same principle, and it will be discussed in a later
chapter.
Chapter 4 Properties of Rock Materials 4

4.2 Mechanical Properties of Rock Material

4.2.1 Compressive Strength

Compressive strength is the capacity of a material to withstand axially directed


compressive forces. The most common measure of compressive strength is the uniaxial
compressive strength or unconfined compressive strength. Usually compressive strength
of rock is defined by the ultimate stress. It is one of the most important mechanical
properties of rock material, used in design, analysis and modelling.

Figure 4.2.1a presents a typical stress-strain curve of a rock under uniaxial compression.
The complete stress-strain curve can be divided into 6 sections, represent 6 stages that the
rock material is undergoing. Figure 4.2.1b and Figure 4.2.1c show the states of rock in
those stages of compression.

Figure 4.2.1a Typical uniaxial compression stress-strain curve of rock material.

Figure 4.2.1b Uniaxial compression test and failure simulated by RFPA.

Figure 4.2.1c Samples of rock material under uniaxial compression test and failure.

Stage I – The rock is initially stressed, pre-existing microcracks or pore orientated at large
angles to the applied stress is closing, in addition to deformation. This causes an initial
non-linearity of the axial stress-strain curve. This initial non-linearity is more obvious in
weaker and more porous rocks,

Stage II – The rock basically has a linearly elastic behaviour with linear stress-strain
curves, both axially and laterally. The Poisson's ratio, particularly in stiffer unconfined
rocks, tends to be low. The rock is primarily undergoing elastic deformation with
minimum cracking inside the material. Micro-cracks are likely initiated at the later
portion of this stage, of about 35-40% peak strength. At this stage, the stress-strain is
largely recoverable, as the there is little permanent damage of the micro-structure of the
rock material.

Stage III – The rock behaves near-linear elastic. The axial stress-strain curve is near-
linear and is nearly recoverable. There is a slight increase in lateral strain due to
dilation. Microcrack propagation occurs in a stable manner during this stage and that
microcracking events occur independently of each other and are distributed throughout
the specimen. The upper boundary of the stage is the point of maximum compaction and
zero volume change and occurs at about 80% peak strength.
Chapter 4 Properties of Rock Materials 5

Stage IV – The rock is undergone a rapid acceleration of microcracking events and


volume increase. The spreading of microcracks is no longer independent and clusters of
cracks in the zones of highest stress tend to coalesce and start to form tensile fractures or
shear planes - depending on the strength of the rock.

Stage V – The rock has passed peak stress, but is still intact, even though the internal
structure is highly disrupt. In this stage the crack arrays fork and coalesce into
macrocracks or fractures. The specimen is undergone strain softening (failure)
deformation, i.e., at peak stress the test specimen starts to become weaker with increasing
strain. Thus further strain will be concentrated on weaker elements of the rock which
have already been subjected to strain. This in turn will lead to zones of concentrated strain
or shear planes.

Stage VI – The rock has essentially parted to form a series of blocks rather than an intact
structure. These blocks slide across each other and the predominant deformation
mechanism is friction between the sliding blocks. Secondary fractures may occur due to
differential shearing. The axial stress or force acting on the specimen tends to fall to a
constant residual strength value, equivalent to the frictional resistance of the sliding
blocks.

In underground excavation, we often are interested in the rock at depth. The rock is
covered by overburden materials, and is subjected to lateral stresses. Compressive
strength with lateral pressures is higher than that without. The compressive strength with
lateral pressures is called triaxial compressive strength. The true triaxial compression
means the 3 different principal stresses in three directions. A true triaxial compression

equal, i.e., σ2 = σ3. In the test, only a confining pressure is required.


testing machine is rather difficult to operate. In most tests, two lateral stresses are made

Figure 4.2.1d shows the results of a series triaxial compression tests. In addition to the
significant increase of strength with confining pressure, the stress-strain characteristics
also changed. Discussion on the influence of confining pressure to the mechanical
characteristics is given in a later section. Typical strengths and modulus of common
rocks are given in Table 4.2.1a.

Figure 4.2.1d Triaxial compression test and failure


Chapter 4 Properties of Rock Materials 6

Table 4.2.1a Mechanical properties of rock materials.


Rock UC Strength Tensile Strength Elastic Modulus Poisson’s Ratio Strain at Failure Point Load Index Fracture Mode I
(MPa) (MPa) (GPa) (%) Is(50) (MPa) Toughness
Igneous
Granite 100 – 300 7 – 25 30 – 70 0.17 0.25 5 – 15 0.11 – 0.41
Dolerite 100 – 350 7 – 30 30 – 100 0.10 – 0.20 0.30 >0.41
Gabbro 150 – 250 7 – 30 40 – 100 0.20 – 0.35 0.30 6 – 15 >0.41
Rhyolite 80 – 160 5 – 10 10 – 50 0.2 – 0.4
Andesite 100 – 300 5 – 15 10 – 70 0.2 10 – 15
Basalt 100 – 350 10 – 30 40 – 80 0.1 – 0.2 0.35 9 – 15 >0.41
Sedimentary
Conglomerate 30 – 230 3 – 10 10 – 90 0.10 – 0.15 0.16
Sandstone 20 – 170 4 – 25 15 – 50 0.14 0.20 1–8 0.027 – 0.041
Shale 5 – 100 2 – 10 5 – 30 0.10 0.027 – 0.041
Mudstone 10 – 100 5 – 30 5 – 70 0.15 0.15 0.1 – 6
Dolomite 20 – 120 6 – 15 30 – 70 0.15 0.17
Limestone 30 – 250 6 – 25 20 – 70 0.30 3–7 0.027 – 0.041
Metamorphic
Gneiss 100 – 250 7 – 20 30 – 80 0.24 0.12 5 – 15 0.11 – 0.41
Schist 70 – 150 4 – 10 5 – 60 0.15 – 0.25 5 – 10 0.005 – 0.027
Phyllite 5 – 150 6 – 20 10 – 85 0.26
Slate 50 – 180 7 – 20 20 – 90 0.20 – 0.30 0.35 1–9 0.027 – 0.041
Marble 50 – 200 7 – 20 30 – 70 0.15 – 0.30 0.40 4 – 12 0.11 – 0.41
Quartzite 150 – 300 5 – 20 50 – 90 0.17 0.20 5 – 15 >0.41
Chapter 4 Properties of Rock Materials 7

4.2.2 Young's Modulus and Poisson’s Ratio

Young's Modulus is modulus of elasticity measuring of the stiffness of a rock material. It


is defined as the ratio, for small strains, of the rate of change of stress with strain. This
can be experimentally determined from the slope of a stress-strain curve obtained during
compressional or tensile tests conducted on a rock sample.

Similar to strength, Young’s Modulus of rock materials varies widely with rock type.
For extremely hard and strong rocks, Young’s Modulus can be as high as 100 GPa.
There is some correlation between compressive strength and Young’s Modulus, and
discussion is given in a later section.

Poisson’s ratio measures the ratio of lateral strain to axial strain, at linearly-elastic region.
For most rocks, the Poisson’s ratio is between 0.15 and 0.4. As seen from early section,
at later stage of loading beyond linearly elastic region, lateral strain increase fast than the
axial strain and hence lead to a higher ratio.

4.2.3 Stress-Strain at and after Peak

With well controlled compression test, a complete stress-strain curve for a rock specimen
can be obtained, as typically shown in Figure 4.2.3a.

Figure 4.2.3a Complete stress-strain curves of several rocks showing post peak
behaviour (Brady and Brown).

Strain at failure is the strain measured at ultimate stress. Rocks generally fail at a small
strain, typically around 0.2 to 0.4% under uniaxial compression. Brittle rocks, typically
crystalline rocks, have low strain at failure, while soft rock, such as shale and mudstone,
could have relatively high strain at failure. Strain at failure sometimes is used as a
measure of brittleness of the rock. Strain at failure increases with increasing confining
pressure under triaxial compression conditions.

Rocks can have brittle or ductile behaviour after peak. Most rocks, including all
crystalline igneous, metamorphic and sedimentary rocks, behave brittle under uniaxial
compression. A few soft rocks, mainly of sedimentary origin, behave ductile.

4.2.4 Tensile Strength

Tensile strength of rock material is normally defined by the ultimate strength in tension,
i.e., maximum tensile stress the rock material can withstand.
Chapter 4 Properties of Rock Materials 8

Rock material generally has a low tensile strength. The low tensile strength is due to the
existence of microcracks in the rock. The existence of microcracks may also be the
cause of rock failing suddenly in tension with a small strain.

Tensile strength of rock materials can be obtained from several types of tensile tests:
direct tensile test, Brazilian test and flexure test. Direct test is not commonly performed
due to the difficulty in sample preparation. The most common tensile strength
determination is by the Brazilian tests. Figure 4.2.4a illustrates the failure mechanism of
the Brazilian tensile tests.

Figure 4.2.4a Stress and failure of Brazilian tensile tests by RFPA simulation.

4.2.5 Shear Strength

Shear strength is used to describe the strength of rock materials, to resist deformation due
to shear stress. Rock resists shear stress by two internal mechanisms, cohesion and
internal friction. Cohesion is a measure of internal bonding of the rock material.

friction angle, φ. Different rocks have different cohesions and different friction angles.
Internal friction is caused by contact between particles, and is defined by the internal

Shear strength of rock material ca be determined by direct shear test and by triaxial
compression tests. In practice, the later methods is widely used and accepted.

With a series of triaxial tests conducted at different confining pressures, peak stresses (σ1)
are obtained at various lateral stresses (σ3). By plotting Mohr circles, the shear envelope
is defined which gives the cohesion and internal friction angle, as shown in Figure 4.2.5a.

Figure 4.2.5a Determination of shear strength by triaxial tests.

Tensile and shear strengths are important as rock fails mostly in tension and in shearing,
even the loading may appears to be compression. Rocks generally have high
compressive strength so failure in pure compression is not common.

4.3 Effects of Confining and Pore Water Pressures on Strength and Deformation

4.3.1 Effects of Confining Pressure

Figure 4.3.1a illustrates a number of important features of the behaviour of rock in triaxial
compression. It shows that with increasing confining pressure,
(a) the peak strength increases;
Chapter 4 Properties of Rock Materials 9

(b) there is a transition from typically brittle to fully ductile behaviour with the
introduction of plastic mechanism of deformation;
(c) the region incorporating the peak of the axial stress-axial strain curve flattens and
widens;
(d) the post-peak drop in stress to the residual strength reduces and disappears at high
confining stress.

Figure 4.3.1a Complete axial stress-axial strain curves obtained in triaxial compression
tests on Tennessee Marble at various confining pressures (after Wawersik & Fairhurst
1970).

The confining pressure that causes the post-peak reduction in strength disappears and the
behaviour becomes fully ductile (48.3 MPa in the figure), is known as the brittle-ductile
transition pressure. This brittle-ductile transition pressure varies with rock type. In
general, igneous and high grade metamorphic rocks, e.g., granite and quartzite, remain
brittle at room temperature at confining pressures of up to 1000 MPa or more. In these
cases, ductile behaviour will not be of concern in practical civil engineering problems.

4.3.2 Effects of Pore Water Pressure

The influence of pore-water pressure on the behaviour of porous rock in the triaxial
compression tests is illustrated by Figure 4.3.2a. A series of triaxial compression tests
was carried out on a limestone with a constant confining pressure of 69 MPa, but with
various level of pore pressure (0-69 MPa). There is a transition from ductile to brittle

response is controlled by the effective confining stress (σ3' = σ3 – u).


behaviour as pore pressure is increased from 0 to 69 MPa. In this case, mechanical

Figure 4.3.2a Effect of pore pressure on the stress-strain behaviour of rock materials.

Effect of pore water pressure is only applicable for porous rocks where sufficient pore
pressure can be developed within the materials. For low porosity rocks, the classical
effective stress law does not hold.

4.3.3 Effects of Intermediate Stress

The axisymmetric triaxial test is a simplified method to determine the axial strength at a

Under the true triaxial compression condition (σ2 > σ3), it is found that σ2 has effects on
confining pressure. The effect of intermediate principal stress (σ2) is not considered.

the axial strength. Figure 4.3.3a shows some typical results of the true triaxial test on
granite and marble (Mogi, Yu).
Chapter 4 Properties of Rock Materials 10

Figure 4.3.3a Change of triaxial compressive strength with intermediate principal


stress.

It is evident that rock triaxial compressive strength increases with σ2 with a fixed σ3.

criterion. It is further suggested that the σ2 effect have two zones: (i) rock triaxial
This is a conclusion quite different from the conventional Mohr-Coulomb strength

compressive strength increases with increasing σ2 from σ2 = σ3 to a critical value (σ2′);


and (ii) rock triaxial compressive strength decreases with further increasing of σ2 from σ2
= σ2′ to σ2 = σ1, as shown in Figure 4.3.3b.

Figure 4.3.3b Effect of intermediate stress on peak strength.

Within the practical range of civil engineering, σ2 and σ3 will generally significant small
compare to the strength. Having σ2 = σ3, tests give a slightly lower and more

the engineering depths could reach a few thousands meters, the effect of σ2 on the
conservative estimate of the strength. However, in petroleum engineering applications,

estimation of strength could be significant in engineering design.

4.4 Other Engineering Properties of Rock Materials

4.4.1 Point Load Strength Index

Point load test is another simple index test for rock material. It gives the standard point
load index, Is(50), calculated from the point load at failure and the size of the specimen,
with size correction to an equivalent core diameter of 50 mm. Description of point load
test and calculation is given later in the section dealing with laboratory test methods.

Typical values of point load strength index for various rocks are given in Table 4.2.1a.

4.4.2 Fracture Toughness

Fracture toughness of rock materials measures the effectiveness of rock fracturing. It is


typically measured by a toughness test. There are three fracture mode: (Mode I), (Mode
II) (Mode III), as shown in Figure 4.4.2a. Correspondingly, there are three fracture
toughness, KIC, KIIC and KIIIC.

Figure 4.4.2a Three typical fracture modes.


Chapter 4 Properties of Rock Materials 11

In rock mechanics, Mode I is associated with the crack initiation and propagation in a
rock material. Table 4.2.1a gives values of KIC for typical rocks.

4.4.3 Brittleness

Brittleness can be expressed by several ways.

4.4.4 Indentation

4.4.5 Swelling

Some rocks swell when they are situated with water. As discussed in Chapter 2,
swelling is governed by the amount of swelling montmorillonite clay minerals in the
rock material.

Rock swelling is measured in confined and unconfined conditions. Unconfined


swelling is measured by the percentage increase of length in three perpendicular
directions, when a rock specimen is placed in water.

Confined swelling index measures swelling in one direction while deformations in other
two directions are constrained.

4.5 Relationships between Physical and Mechanical Properties

4.5.1 Rock Hardness, Density, and Strength

Schmidt hammer rebound hardness is often measured during early part of field
investigation. It is a measure of the hardness of the rock material by count the rebound
degree. At the same time, the hardness index can be used to estimate uniaxial
compressive strength of the rock material. The correlation between hardness and
strength is shown in Figure 4.5.1a. The correlation is also influenced by the density of
the material.

Figure 4.5.1a Correlation between hardness, density and strength.

4.5.2 Effect of Water Content on Strength


Chapter 4 Properties of Rock Materials 12

Many tests showed that the when rock materials are saturated or in wet condition, the
uniaxial compressive strength is reduced, compared to the strength in dry condition.

4.5.3 Velocity and Modulus

While seismic wave velocity gives a physical measurement of the rock material, it is also
used to estimate the elastic modulus of the rock material. From the theory of elasticity,
compressional (or longitudinal) P-wave velocity (vp) is related to the elastic modulus (Es),
and the density (ρ) of the material as,

vp = (Es / ρ)½ or Es = ρ vp2

If ρ in g/cm3, and vp in km/s, then Es in GPa (109 N/m2). The elastic modulus estimated
by this method is the sometime termed as seismic modulus (also called dynamic modulus,
but should not be mistaken as the modulus under dynamic compression). It is different
from the modules obtained by the uniaxial compression tests. The value of the seismic
modulus is generally slightly higher than the modulus determined from static compression
tests.

Similarly, seismic shear modulus Gs may be determined from shear S-wave velocity vs,

Gs = ρ vs2

Gs is in GPa, when density ρ is in g/cm3, and S-wave velocity vs is in km/s.


Poisson’s ration νs can be determined from,
Seismic

νs =
1 – 2 (vs / vp)2
2 [1 – (vs / vp)2]

Alternatively, seismic Young’s modulus Es can be determined from shear modulus (Gs)
and Poisson’s ratio (νs),

Es = 2 Gs (1 + νs)

4.5.4 Compressive Strength and Modulus

It is a general trend that a stronger rock material is also stiffer, i.e., higher elastic modulus
is often associated with higher strength. There is reasonable correlation between
compressive strength and elastic modulus. The correlations are presented in Figure
4.5.4a.

Figure 4.5.4a Correlation between strength and modulus.


Chapter 4 Properties of Rock Materials 13

It should be noted that the correlation is not precisely linear and also depends on the rock
type, or perhaps on the texture of the rocks.

4.5.5 Compressive and Tensile Strengths

As shown by the Griffith criterion, tensile strength of brittle materials is theoretical 1/8 of
the compressive strength. Typically, tensile strength of rock materials is about 1/10 to
1/8 of the compressive strength.

It is important to be aware that compressive strength is significantly greater than tensile


strength for rocks. Therefore, rock fails easily under tension. In design, rock should be
subjected to minimum tensile stress.

4.5.6 Point Load Index and Strengths

With massive tests results, correlations between the point load index (Is(50))and uniaxial
compressive strength (σc)and between the point load index and Brazilian tensile strength
(σt)are suggested as below,

σc ≈ 22 Is(50)
σt ≈ 1.25 Is(50)

However, it is noted the correlation factor between σc and Is(50) can vary between 10 to 30.
By theoretical analysis, the correlation factor is 12.5 (Chua and Wong 1996). It is
strongly suggested that the point load index should be used as an independent strength
index, and should not used to determine the compressive strength.

The correlation between σt and Is(50) is more consistent, and again, it can vary with a
significant margin. The failure mode for point load test is primarily by tensile
fracturing.

4.5.7 Point Load Index and Fracture Toughness

As the point load failure is due to Mode I fracturing, it is therefore anticipated some
correlations between Mode I Fracture Toughness (KIC) and point load index (Is(50)). Two
correlations below are suggested based on test results,

KIC ≈ 1.1 + 0.1 Is(50)

KIC ≈ 0.21 Is(50)


Chapter 4 Properties of Rock Materials 14

those of the second equation. When Is(50) ≈10 (typically for crystalline rocks), both
The first equation gives a higher estimate at Is(50)<10, and lower estimate at Is(50)>10, than

correlation equations give KIC ≈ 2.1.

4.6 Failure Criteria of Rock Materials

4.6.1 Mohr-Coulomb criterion

Mohr-Coulomb strength criterion assumes that a shear failure plane is developed in the
rock material. When failure occurs, the stresses developed on the failure plane are on

normal stress σn and shear stress τ.


the strength envelope. Refer to Figure 4.6.1a, the stresses on the failure plane a-b are the

Figure 4.6.1a Stresses on failure plane a-b and representation of Mohr’s circle.

Applying the stress transformation equations or from the Mohr’s circle, it gives:

σn = ½ (σ1 + σ3) + ½ (σ1 – σ3) cos2θ

τ = ½ (σ1 – σ3) sin2θ

Coulomb suggested that shear strengths of rock are made up of two parts, a constant
cohesion (c) and a normal stress-dependent frictional component, i.e.,

τ = c + σn tanφ

where c = cohesion and φ = angle of internal friction.

Therefore, by combining the above three equations,

½ (σ1 – σ3) sin2θ = c + [½ (σ1 + σ3) + ½ (σ1 – σ3) cos2θ] tanφ

or

2c + σ3 [sin2θ + tanφ (1 – cos2θ)]


σ1 =
sin2θ – tanφ (1 + cos2θ)

In a shear stress-normal stress plot, the Coulomb shear strength criterion τ = c + σn tanφ is
represented by a straight line, with an intercept c on the τ axis and an angle of φ with the
σn axis. This straight line is often called the strength envelope. Any stress condition
below the strength envelope is safe, and once the stress condition meet the envelope,
failure will occur.
Chapter 4 Properties of Rock Materials 15

As assumed, rock failure starts with the formation of the shear failure plane a-b.
Therefore, the stress condition on the a-b plane satisfies the shear strength condition. In
another word, the Mohr-Coulomb strength envelope straight line touches (makes a
tangent) to the Mohr’s circles. At each tangent point, the stress condition on the a-b
plane meets the strength envelope.

As seen from the Mohr’s circle, the failure plane is defined by θ, and

θ=¼π+½φ

Then

2c cosφ + σ3 (1 + sinφ)]
σ1 =
1 – sinφ

Figure 4.6.1b Mohr-Coulomb strength envelope in terms of normal and shear stresses
and principal stresses, with tensile cut-off.

uniaxial compressive strength is related to c and φ by:


If the Mohr-Coulomb strength envelope shown in Figure 4.6.1b is extrapolated, the

σc =
2 c cosφ
1 – sinφ

An apparent value of uniaxial tensile strength of the material is given by:

σt =
2 c cosφ
1 + sinφ

However, the measured values of tensile strength are generally lower than those predicted

value of uniaxial tensile stress, σt′, as shown in Figure 4.6.1b. For most rocks, σt′ is
by the above equation. For this reason, a tensile cut-off is usually applied at a selected

about 1/10 σc.

The Mohr-Coulomb strength envelope can also be shown in σ1–σ3 plots, as seen in Figure
4.6.1b. Then,

σ1 = σc + σ3 tanψ

and

1 + sinφ
tanψ =
1 – sinφ
Chapter 4 Properties of Rock Materials 16

or

σ1 = σc + σ3
1 + sinφ
1 – sinφ

The Mohr-Coulomb criterion is only suitable for the low range of σ3. At high σ3, it

engineering deals with shallow problems and low σ3, so the criterion is widely used, due
overestimates the strength. It also overestimates tensile strength. In most cases, rock

to its simplicity and popularity.

4.6.2 Griffith strength criterion

Based on the energy instability concept, Griffith extended the theory to the case of
applied compressive stresses. Assuming that the elliptical crack will propagate from the
points of maximum tensile stress concentration (P in Figure 4.6.2a), Griffith obtained the
following criterion for crack extension in plane compression:

Figure 4.6.2a Griffith crack model for plane compression.

(σ1 – σ3)2 – 8 σt (σ1 + σ3) = 0 if σ1 + 3 σ3 > 0

σ1 + σt = 0 if σ1 + 3 σ3 < 0

where σt is the uniaxial tensile strength of the material.

When σ3 = 0, the above equation becomes

σ1 – 8 σt = 0

It in fact suggests that the uniaxial compressive stress at crack extension is always eight
times the uniaxial tensile strength

Figure 4.6.2b Griffith envelope for crack extension in compression.

The strength envelopes given by the above equations in principal stresses and in normal
and shear stresses are shown in Figure 4.6.2b. This criterion can also be expressed in
terms of the shear stress (τ) and normal stress (σn) acting on the plane containing the
major axis of the crack:

τ2 = 4 σt (σn + σt)

When σn = 0, τ = 2σt, which represents the cohesion.


Chapter 4 Properties of Rock Materials 17

The classic plane compression Griffith theory did not provide a very good model for the
peak strength of rock under multiaxial compression. It gives only good estimate of
tensile strength, and under-estimate compressive strength. Accordingly, a number of
modifications to Griffith's solution were introduced. These criteria do not find practical
use today.

4.6.3 Hoek-Brown criterion

Because the classic strength theories used for other engineering materials have been found
not to apply to rock over a wide range of applied compressive stress conditions, a number
of empirical strength criteria have been introduced for practical use. One of the most
widely used criteria is Hoek-Brown criterion for isotropic rock materials and rock masses.

Hoek and Brown (1980) found that the peak triaxial compressive strengths of a wide
range of isotropic rock materials could be described by the following equation:

σ1 σ σ
σc σc σc
= 3 + ( m 3 + 1.0)½

or

σ1 = σ3 + (m σ3 σc + σc2)½

Where m is a parameter that changes with rock type in the following general way:

(a) m ≈ 7 for carbonate rocks with well developed crystal cleavage (dolomite, limestone,

(b) m ≈ 10 for lithified argillaceous rocks (mudstone, siltstone, shale, slate);


marble);

(c) m ≈ 15 for arenaceous rocks with strong crystals and poorly developed crystal

(d) m ≈ 17 for fine-grained polyminerallic igneous crystalline rocks (andesite, dolerite,


cleavage (sandstaone, quartzite);

(e) m ≈ 25 for coarse-grained polyminerallic igneous and metamorphic rocks


diabase, rhyolite);

(amphibolite, gabbro, gneiss, granite, norite, quartz-diorite).

Figure 4.6.3a shows normalized Hoek-Brown peak strength envelope for some rocks. It
is evident that the Hoek-Brown strength envelope is not a straight line, but a curve. At
high stress level, the envelope curves down, so it gives low strength estimate than the
Mohr-Coulomb envelope.

Figure 4.6.3a Normalized peak strength envelope for (i) granites and (ii) sandstones
(after Hoek & Brown 1980).
Chapter 4 Properties of Rock Materials 18

The Hoek-Brown peak strength criterion is an empirical criterion based on substantial test
results on various rocks. It is however very easy to use and select parameters. It is also
extended to rock masses with the same equation, hence makes it is so far the only
acceptable criterion for both material and mass. The Hoek Brown rock mass strength
criterion is covered in a later chapter.

4.7 Effects of Rock Microstructures on Mechanical Properties

4.7.1 Strength of rock material with Anisotropy

Rocks, such as shale and slate, are not isotropic. Because of some preferred orientation
of fabric or microstructure, or the presence of bedding or cleavage planes, the behaviour
of those rocks is anisotropic. There are several forms of anisotropy with various degrees
of complexity. It is therefore only the simplest form of anisotropy, transverse isotropy,
to be discussed here.

The peak strengths developed by transversely isotropic rocks in triaxial compression vary
with the orientation of the plane of isotropy, plane of weakness or foliation plane, with
respect to the principal stress directions. Figure 4.7.1a shows some measured variations
in peak principal stress difference with the angle of inclination of the major principal
stress to the plane of weakness.

Figure 4.7.1a Variation of differential stresses with the inclination angle of the plane of
weakness (see Brady & Brown 1985).

Analytical solution shows that principal stress difference (σ1–σ3) of a transversely


isotropic specimen under triaxial compression shown in Figure 4.7.1a can be given by the
equation below (Brady & Brown 1985):

2 (cw + σ3 tanφw)
(σ1 – σ3) =
(1 – tanφw cotβ) sin2β

φw = angle of friction of the plane;


where cw = cohesion of the plane of weakness;

β = inclination of the plane.

The minimum strength occurs when

tan2β = –cot φw, or β = 45° + ½ φw

The corresponding value of principal stress difference is,


Chapter 4 Properties of Rock Materials 19

(σ1 – σ3)min = 2 (cw + σ3 tanφw) [(1 + tan φw) + tanφw]


2 ½

Figure 4.7.1b shows variation of σ1 at constant σ3 with angle β, plotted using the above
equation. When the weakness plane is at an angle of 45° + ½ φw, the strength is the
lowest. For rocks, φw is about 30° to 50°, hence β is about 60° to 70°. In compression
tests, intact rock specimens generally fails to form a shear plane at an angle about 60° to
70°. This in fact shows that when the rock containing an existing weakness plane that is
about to become a failure plane, the rock has the lowest strength.

Figure 4.7.1b Variation of σ1 at constant σ3 with angle β.

4.7.2 Rocks containing Pores and Bridged Cracks

4.7.3 Effects of Mineral Bounding on Rock Material Strength

4.7.4 Effects of Rock Texture Inhomogeneity on Strength and Failure

4.7.5 Soft and Weathered Rocks

Rocks whose uniaxial compressive strength falls approximately in the range 0.5-25 MPa
as suggested by ISRM is considered as soft rocks or weak rocks. Soft rocks are usually
sediments in the process of consolidation and solidification. Soft or weak rocks can also
be weathered rocks.

Soft rocks may have the isotropic compression response similar in detail to the response
of consolidated soils, as presented in Figure 4.7.5a. Reader should refer to a soil
mechanics textbook for details regarding compressibility.

Figure 4.7.5a Compression characteristics of a soil and a mudstone (Johnston 1993).

The results obtained on over-consolidated London clay showed that for low confining
pressures, the stress-strain behaviour is relative brittle. As the confining pressure was
increased, the response became more ductile. Exactly the same responses are displayed
for soft rocks, and indeed for hard rocks.
Chapter 4 Properties of Rock Materials 20

The behaviour of soft rocks is somewhere between soil and hard rock. For soil, strength
criterion is generally represented by the Mohr-Coulomb criterion, while for hard rocks, it
is by the Hoek-Brown criterion. The former is a straight line and the later is a curve.

The strength characteristics of soft rocks are somehow, between soil and hard rock and
are described by parabolic curves shown in Figure 4.7.5b.

Figure 4.7.5b Strength envelopes of soils, soft rocks and hard rocks (Johnston 1993).

With increase of strength of geomaterials, there is a general progression from the linear
strength criterion of Mohr-Coulomb for soft soil to the curved strength criterion of
Hoek-Brown for competent rocks. A criterion describing this progression is given by
Johnston (1993) as,

σ1 M σ3
σc B σc
B
=( + 1) (4.7.5a)

where

1 + sin φ
1 – sin φ
M=

The criterion only considers effect strength and hence φ is the drained friction angle. B is
a strength dependent constant, 1.0 ≥ B ≥ 0.5. For soft soil, B approaches to 1.0, the
above equation becomes the Mohr-Coulomb criterion as below,

σ1 σ 1 + sin φ
σc σc 1 – sin φ
=1+ 3 (4.7.5b)

For rocks, as the strength increases, M increases, but B reduces approaching to 0.5.
The equation then becomes a parabolic envelope very similar to that of the Hoek-Brown
criterion.

4.8 Time Dependent Characteristics of Rock Materials

4.8.1 Rheologic Properties of Rock Materials

4.8.2 Effect of Loading Rate on Rock Strength


Chapter 4 Properties of Rock Materials 21

4.8.3 Failure Mechanism of Rock Material under Impact and Shock Loading

4.9 Laboratory Testing of Rock Materials

This section provides method statement of the common rock mechanics tests.
International Society for Rock Mechanics (ISRM) suggested methods is freely available
from the ISRM website (www.isrm.net). Readers should refer to the ISRM methods for
details.

4.9.1 Petrographic Description

Thin section is prepared and examined under petrographic microscope equipped with
modal analysis device, e.g., point counter from a grid over the thin section, as shown in
Figure 4.9.1a.

Figure 4.9.1a Petrographic analysis.

Petrographic analysis is reported with the form below.

Project: Location:
Co-ordinates: Specimen No.:
Description of sampling point:
Thin section No.: Date:
GEOLOGICAL DESCRIPTION MACROSCOPIC DESCRIPTION
Rock name: Degree of weathering:
Petrographic classification: Structure (incl. bedding):
Geological formation: Discontinuities:
MINERAL COMPOSITION GRAIN SIZE
DISTRIBUTION
Major Vol. Minor Vol. Accessory Vol. Microns %
Component % Component % %

QUALITATIVE DESCRIPTION SIGNIFICANCE FOR ROCK ENGINEERING


Texture:
Fracturing:
Alteration:
Matrix:
GENERAL REMARK MICROGRAPH PHOTO OF TYPICAL
FEATURES OF THIN SECTION
Chapter 4 Properties of Rock Materials 22

International Society for Rock Mechanics (ISRM), Commission on Testing Methods


(1978), Suggested methods for petrographic description of rocks.

4.9.2 Density, Porosity and Water Content

(a) Density of Samples of Regular Shape

The mass of the samples is weighed and its volume is calculated. The bulk density is the
mass per unit volume.

(b) Density of Sample of Irregular Shape

The mass of the samples is weighed (Mbulk), the rock samples are saturated by water
immersion in a vacuum of 600 Pa for more than 1 hour, with periodic agitation to
remove trapped air. The samples are then transferred underwater to a basket in an
immersion bath. Their saturated-submerged mass Msub is determined from the
difference between the saturated-submerged mass of the basket plus sample and that of
the basket alone. The samples are then taken out from the immersion bath and surface
dried with a moist cloth, their saturated-surface-dry mass Msat is recorded. The bulk
volume (Vbulk) and bulk density (ρbulk) are calculated as:

ρw
(Msat – Msub)
Vbulk =

ρbulk =
Mbulk
Vbulk

(c) Porosity of Rock Samples

Mbulk, Msat, Msub and Vbulk are determined using the same method as that for density
measurement of rock samples with irregular shape. The samples are dried to a constant
mass at a temperature of 105°C in an oven and cooled for 30 min in a desiccator, and the
dry mass Mdry is measured. Pore volume (Vv) and porosity (n) are calculated by

ρw
(Msat – Mdry)
Vv =

× 100%
Vv
n=
Vbulk
Chapter 4 Properties of Rock Materials 23

(d) Water Content of Rock Samples

Mbulk, and Mdry are determined using the same method as that for density and porosity
measurement of rock samples with irregular shape. Water content (w) is calculated by

× 100%
(Mbulk – Mdry)
w=
Mdry

Reporting of results includes description of the rocks, precautions taken to retain water
during sampling and storage.

International Society for Rock Mechanics (ISRM), Commission on Testing Methods


(1977), Suggested methods for determining water content, porosity, density, absorption
and related properties and swelling and slake-durability index properties.

4.9.3 Compression Tests

(a) Uniaxial Compression Strength Test

Specimens of right circular cylinders having a height to diameter ratio of 2 or higher are
prepared by cutting and grinding. Two axial and one circumferential deformation
measurement devices (LVDTs) are attached to each of the specimen. The specimen is

stress is applied with a constant strain rate around 1 μm/s such that failure occurs within
then compressed under a stiff compression machine with a spherical seating. The axial

5-10 minutes of loading. The load is measured by a load transducer. Load, two axial
deformations and one circumferential deformation measurements are recorded at every 2-
5 KN interval until failure. Uniaxial compressive strength, Young's modules (at 50% of
failure stress) and Poisson's ratio (at 50% of failure stress) can be calculated from the
failure load, stress and strain relationship.

Uniaxial compressive strength, σc, is calculated as the failure load divided by the initial
cross sectional area of the specimen.

Axial tangential Young's modulus at 50% of uniaxial compressive strength, Et50% is


calculated as the slope of tangent line of axial stress - axial strain curve at a stress level
equals to 50% of the ultimate uniaxial compressive strength.

Poisson's ratio at 50% of uniaxial compressive strength, ν50%, is calculated as:

slope of axial stress-strain curve at 50% of σc


ν50% = –
slope of lateral stress-strain curve at 50% of σc

Reporting of results includes description of the rock, specimen anisotropy, specimen


dimension, density and water content at time of test, mode of failure, uniaxial
Chapter 4 Properties of Rock Materials 24

compressive strength, modulus of elasticity, Poisson's ratio, stress-strain (axial and lateral)
curves to failure.

Figure 4.9.3a A typical uniaxial compression test set-up with load and strain
measurements.

(b) Triaxial Compression Strength Test

Specimens of right circular cylinders having a height to diameter ratio of 2 or higher are
prepared by cutting and grinding. Two axial and two lateral deformation (or a
circumferential deformation if a circumferential chain LVDT device is used),
measurement devices are attached to each of the specimen. The specimen is placed in a
triaxial cell (e.g., Hoek-Franklin cell) and a desired confining stress is applied and
maintained by a hydraulic pump. The specimen is then further compressed under a stiff

strain rate around 1 μm/s such that failure occurs within 5-15 minutes of loading. The
compression machine with a spherical seating. The axial stress is applied with a constant

load is measured by a load transducer. Load, 2 axial strain or deformation and 2 lateral
strains or deformation (or a circumferential deformation if a circumferential chain LVDT
device is used) are recorded at a fixed interval until failure. Triaxial compressive
strength, Young's modules (at 50% of failure stress) and Poisson's ratio (at 50% of failure
stress) can be calculated from the axial failure load, stress and strain relationship.

Triaxial compressive strength, σ1, is calculated as the axial failure load divided by the
initial cross sectional area of the specimen.

Axial tangential Young's modulus at 50% of triaxial compressive strength, Et50% is


calculated as the slope of tangent line of axial stress - axial strain curve at a stress level
equals to 50% of the ultimate uniaxial compressive strength.

Poisson's ratio at 50% of triaxial compressive strength is calculated with the same
methods as for the uniaxial compression test.

circle are plotted using confining stress as σ3 and axial stress as σ1. Failure envelopes
For a group of triaxial compression tests at different confining stress level, Mohr's stress

(Mohr, Coulomb or Hoek and Brown) and parameters of specified failure criterion are
determined.

Reporting of results includes description of the rock, specimen anisotropy, specimen


dimension, density and water content at time of test, mode of failure, triaxial compressive
strength, modulus of elasticity, Poisson's ratio, stress-strain (axial and lateral) curves to
failure, Mohr's circles and failure envelope.

Figure 4.9.3b Triaxial compression test using Hoek cell.


Chapter 4 Properties of Rock Materials 25

International Society for Rock Mechanics (ISRM), Commission on Testing Methods


(1979), Suggested methods for determining the uniaxial compressive strength and
deformability of rock materials.

International Society for Rock Mechanics (ISRM), Commission on Testing Methods


(1999), Suggested methods for complete stress-strain curve for intact rock in uniaxial
compression.

International Society for Rock Mechanics (ISRM), Commission on Testing Methods


(1978), Suggested methods for determining the strength of rock materials in triaxial
compression.

International Society for Rock Mechanics (ISRM), Commission on Testing Methods


(1983), Suggested methods for determining the strength of rock materials in triaxial
compression: revised version.

4.9.4 Tensile Tests

(a) Direct Tension Test

Direct tension tests on rock materials are not common, due to the difficulty in specimen
preparation. For direct tension test, rock specimen is to be prepared in dog-bone shape
with a thin middle. The specimen is then loaded in tension by pulling from the two
ends.

Deformation modulus can be measured by having strain gauges attached to the specimen.
calculation and the Young’s modulus and the Poisson’s ratio is similar to that for the
uniaxial compression test.

(b) Brazilian Tensile Strength Test

Cylindrical specimen of diameter approximately equals to 50 mm and thickness


approximately equal to the radius is prepared. The cylindrical surfaces should be free
from obvious tool marks and any irregularities across the thickness. End faces shall be
flat to within 0.25 mm and square and parallel to within 0.25°. The specimen is wrapped
around its periphery with one layer of the masking tape and loaded into the Brazil tensile
test apparatus across its diameter. Loading is applied continuously at a constant rate
such that failure occurs within 15-30 seconds. Ten specimens of the same sample shall
be tested.

The tensile strength of the rock is calculated from failure load (P), specimen diameter (D)
and specimen thickness (t) by the following formula:

σt = –
0.636 P
Chapter 4 Properties of Rock Materials 26

Dt

Reporting of results includes description of the rock, orientation of the axis of loading
with respect to specimen anisotropy, water content and degree of saturation, test duration
and loading rate, mode of failure.

Figure 4.9.4b Brazilian tensile test.

(c) Flexure Tension Test

International Society for Rock Mechanics (ISRM), Commission on Testing Methods


(1978), Suggested methods for determining tensile strength of rock materials.

4.9.5 Shear Strength Tests

(a) Direct Punch Shear

(b) Shear Strength Determination by Triaxial Compression Results

Shear strength parameters, cohesion (c) and international friction angle (φ) can be
determined from triaxial compression test data.

The Mohr’s circle can be plotted for a series of triaxial tests results with σ1 at different σ3,
forming a series circles, as typically shown in the figure below. A straight line is draw

envelope. The angle of the line to the horizontal is the internal friction angle φ, and the
to fit best by tangent to all the Mohr’s circles. The line represents the shear strength

intercept at τ axis is the cohesion c.

Alternatively, a series equation can be formed for sets of σ1 and σ1, based on the Mohr-
Coulomb criterion,

2c cosφ + σ3 (1 + sinφ)]
σ1 =
1 – sinφ

Cohesion c and friction angle φ can be computed by solving the equations.

4.9.6 Point Load Strength Index Test

Point load test of rock cores can be conducted diametrically and axially. In diametrical
test, rock core specimen of diameter D is loaded between the point load apparatus across
its diameter. The length/diameter ratio for the diametrical test should be greater than
Chapter 4 Properties of Rock Materials 27

1.0. For axial test, rock core is cut to a height between 0.5 D to D and is loaded between
the point load apparatus axially. Load at failure is recorded as P. Uncorrected point
load strength, Is, is calculated as:

I s = P / D e2

where De, the "equivalent core diameter", is given by:

D e2 = D 2 for diametrical test;


= 4 A /π for axial, block and lump tests;

A = H D = minimum cross sectional area of a plane through the loading points.

of 50 mm. For De ≠ 50 mm, the size correction factor is:


The point load strength is corrected to the point load strength at equivalent core diameter

0.45
F = (De / 50)

The corrected point load strength index, Is(50) is calculated as:

Is(50) = F Is

Figure 4.9.6a Point load test.

International Society for Rock Mechanics (ISRM), Commission on Testing Methods


(1985), Suggested method for determining point load strength.

4.9.7 Ultrasonic wave velocity

Cylindrical rock sample is prepared by cutting and lapping the ends. The length is
measured. An ultrasonic digital indicator consist a pulse generator unit, transmitter and
receiver transducers are used for sonic pulse velocity measurement. The transmitter and
the receiver are positioned at the ends of specimen and the pulse wave travel time is
measured. The velocity is calculated from dividing the length of rock sample by wave
travel time.

Both P-wave and S-wave velocities can be measured.

Figure 4.9.7a Measuring P and S wave velocity in a rock specimen.

International Society for Rock Mechanics (ISRM), Commission on Testing Methods


(1978), Suggested methods for determining sound velocity.
Chapter 4 Properties of Rock Materials 28

4.9.8 Hardness

(a) Schmidt Hammer Rebound Hardness

A Schmidt hammer with rebound measurement is used for this test. The Schmidt hammer
is point perpendicularly and touch the surface of rock. The hammer is released and
reading on the hammer is taken. The reading gives directly the Schmidt hammer
hardness value. The standard Schmidt hardness number is taken when the hammer is
point vertically down. If the hammer is point to horizontal and upward, correction is
needed to add to the number from the hammer. The correction number is given in Table
4.9.8a.

Table 4.9.8a Correction number for Schmidt hammer with inclination angle
Rebound Vertically 45° downward Horizontal 45° upward Vertically
Number downward upward
10 0 –0.8 –3.2
20 0 –0.9 –8.8 –6.9 –3.4
30 0 –0.8 –7.8 –6.2 –3.1
40 0 –0.7 –6.6 –5.3 –2.7
50 0 –0.6 –5.3 –4.3 –2.2
60 0 –0.4 –4.0 –3.3 –1.7

At least 20 tests should be conducted on any one rock specimen. It is suggest to omit 2
lowest and 2 highest reading, and to use the remaining reading for calculating the average
hardness value.

Report of results should include descriptions of rock type, location, size and shape, and
orientation of hammer axis.

Figure 4.9.8a Schmidt hammer rebound hardness test.

(b) Shore Hardness

International Society for Rock Mechanics (ISRM), Commission on Testing Methods


(1978), Suggested methods for determining hardness and abrasiveness of rocks.

International Society for Rock Mechanics (ISRM), Commission on Testing Methods


(2006), Suggested methods for determining the shore hardness value for rock.
Chapter 4 Properties of Rock Materials 29

4.9.9 Fracture Toughness

(a) Cracked Cheron Notched Brazilian Disc (CCNBD)

Figure 4.9.9a CCNBD fracture toughness test set-up-

(b) Chevron Bend Specimen

(c) Short Rod Specimen

International Society for Rock Mechanics (ISRM), Commission on Testing Methods


(1988), Suggested methods for determining the fracture toughness of rock.

International Society for Rock Mechanics (ISRM), Commission on Testing Methods


(1995), Suggested methods for determining Mode I fracture toughness using cracked
Cheron notched Brazilian disc (CCNBD) specimens.

4.9.10 Abrasivity

(a) Cerchar Abrasivity Test

The Cerchar abrasivity test is an abrasive wear with pressure test . It was proposed by
the Laboratoire du Centre d’Etudes et Recherches des Charbonnages (Cerchar) in France.
The testing apparatus is featured in Figure 4.9.10a. It consists of a vice for holding rock
sample (1), which can be moved across the base of the apparatus by a handwheel (2), that
drives a screwthread of pitch 1 mm/revolution turning. Displacement of the vice (1) is
measured by a scale (3). A steel stylus (4), fitting into a holder (5), loaded on the
surface of the rock sample. A dead weight (6) of 70 N is applied on the stylus.

Figure 4.9.10a Cerchar abrasivity test West apparatus (West 1989).

To determine the CAI value the rock is slowly displaced by 10 mm with a velocity of
approximately 1 mm/s. The abrasiveness of the rock is then obtained by measuring the
resulting wear flat on the tip of the steel stylus. The CAI value is calculated as,

CAI = 10-2 d

where d is the wear flat diameter of the stylus tip in μm.

(b) NTNU Abrasion Test


Chapter 4 Properties of Rock Materials 30

West G (1989), Rock abrasiveness testing for tunnelling. International Journal of Rock
Mechanics and Mining Sciences, Vol.26, pp.151-160.

4.9.11 Indentation Test

4.9.12 Slake Durability Test

Select representative rock sample consisting of 10 lumps each of 40-60g, roughly


spherical in shape with corners rounded during preparation. The sample is placed in the
test drum of 2 mm standard mesh cylinder of 100 mm long and 140 mm in diameter with
solid removable lid and fixed base, and is dried to a constant mass at 105°C. The mass
of drum and sample is recorded (Mass A). The sample and drum is placed in trough
which is filled with slaking fluid, usually tap water at 20°C, to a level 20 mm below the
drum axis, and the drum is rotated at 20 rpm for 10 minutes (Figure 4.9.12a). The drum
and sample are removed from trough and oven dried to a constant mass at 105°C without
the lid. The mass of the drum and sample is recorded after cooling (Mass B). The
slaking and drying process is repeated and the mass of the drum and sample is recorded
(Mass C). The drum is brushed clean and its mass is recorded (Mass D).

Figure 4.9.12a Slake durability test.

The slake-durability index is taken as the percentage ratio of final to initial dry sample
masses after to cycles,

× 100%
C–D
Slake-durability index, Id2 =
A–D

The first cycle slake-durability index should be calculated when Id2 is 0-10%,

× 100%
B–D
Slake-durability index, Id1 =
A–D

Table 4.9.12a Slake Durability Classification


Slake Durability Index Id2 (%) Classification
0 – 25 Very low
25 – 50 Low
50 – 75 Medium
75 – 90 High
90 – 95 Very high
95 – 100 Extremely high
Chapter 4 Properties of Rock Materials 31

International Society for Rock Mechanics (ISRM), Commission on Testing Methods


(1977), Suggested methods for determining water content, porosity, density, absorption
and related properties and swelling and slake-durability index properties.

4.9.13 Swelling

(a) Unconfined Swelling Test

Prepared rock specimen of 30-50 mm high is placed in a swelling cell (Figure 4.9.13a).
Micrometer dial gauges reading to 0.002 mm, mounted to measure the swelling
displacements of the specimen in three perpendicular directions. The cell is then flooded
with water to cover the specimen, and the swelling displacements recorded as a function
of time elapsed, until they reach constant levels or pass peaks. The swelling strains are
calculated as

εswell = (maximum swelling displacement / initial length) × 100%

in their respective directions.

Figure 4.9.13a Unconfined swelling test.

(b) Confined Swelling Test

Figure 4.9.13b Confined axial swelling test.

International Society for Rock Mechanics (ISRM), Commission on Testing Methods


(1977), Suggested methods for determining water content, porosity, density, absorption
and related properties and swelling and slake-durability index properties.

Further Readings

Jaeger JC, Cook NGW, Fundamentals of Rock Mechanics, 3rd Edition. Chapman and
Hall, London, 1979.

Brady BHG, Brown ET, Rock Mechanics for Mining Engineers, 2nd Edition.
σ
IV
V
III

VI
II

I
ε

4.2.1a

4.2.1b

1
Ar ou n d pe a k st r e ss ( I V- V) Post pe a k st r e ss ( VI )

4.2.1c

σ1
σ3 increases

σ3 = 0

4.2.1d

2
4.2.3a

4.2.4a

3
σ1 σ3 = 10
τ φ

σ3 = 0
c

σt 0 σ3 σc σ1 σn

σ3 = 0

4.2.5a

4.3.1a

4
σ1
σ1
u= 0

σ3
u

u increases
u = σ3

4.3.2a

4.3.3a

5
Constant σ3 = 30 MPa

4.3.3b

Mode I Mode II Mode III

4.4.2a

6
4.5.1a

160

80 Metamorphic rocks

I gneous rocks
Young’s Modulus, GPa

40
t io
ra
s
u lu
20 od
m
g h
Hi
Sedimentary rocks
10
t io
1 ra t io
0: s ra
lu
50 u us
5 od u l
m od
m :1 m
iu 0 w
ed 20 Lo
M
2.5

6.3 12.5 25 50 100 200 400


Uniaxial Compressive Strength, MPa

4.5.4a

7
τ
φ
σ1

(σn, τ)
b
σn
σ3 2θ
c
τ
σ3 σ1 σn
θ
a

4.6.1a

τ
φ

σt σt′ 0 σ3 σc σ1 σn

σ1

σc

σt σt′ σ3
4.6.1b

8
σ1

Crack extension
σ2
P

σ2

Pre-existing crack
P

σ1

4.6.2a

σ1

8σt (σ1 – σ2)2 =


8 σt (σ1 + σ3)

σt σ2

2σt τ2 = 4 σt (σn + σt)

σt 0 σn

4.6.2b

9
4.6.3a

4.7.1a

10
σ1 σ1

fail of rock
material

σ3
slip on plane
of weakness

45+½φw 90° β

4.7.1b

Typical soil

4.7.5a

11
4.7.5b

4.9.1a

12
4.9.3a

1 1

2
2

3 3

1. Axial load measured by a load cell.


2. Axial displacement or strain
measured by LVDTs or strain
gauges.
3. Circumferential displacement or
lateral strain measured by LVDT or
strain gauges.
4.9.3a

13
4.9.3b

4.9.4b

14
4.9.6a

Transmitter Receiver

Pulse Generator
and Logger

4.9.7a

15
4.9.8a

4.9.9a

16
(1) Rock sample holder.
(2) Hand wheel for
base movement.
(3) Displacement scale.
(4) Steel stylus.
(5) Stylus holder.
(6) Dead weight.

4.9.10a

4.9.12a

17
4.9.13a

(1) Steel ring.


(2) Porous metal plate.
(3) Steel loading plate.
(4) Container.
(5) Dial gauges.
(6) Load measuring device.
(7) Frame.
(8) Loading piston.
(9) Steel plate.

4.9.13b

18
Chapter 5 Properties of Rock Discontinuities 1

CHAPTER 5
PROPERTIES OF ROCK DISCONTINUITIES
Properties of rock discontinuities govern the overall behaviour of the rock masses. This
Chapter addresses properties of rock discontinuities.

Rock discontinuities include joints, fractures, faults and other geological structures.
Rock joints are by far the most common discontinuity encountered in rock masses. Rock
fractures are random features. Rock faults and folds are major but localised geological
structures and therefore are dealt individually.

5.1 Geometrical Characteristics of Rock Joints

5.1.1 Joint Sets and Length: Joints and Fractures, Set Number, and Persistence

As discussed early in the chapter dealing with rock formation, joints are generally in sets,
i.e., parallel joints. The number of joint sets can vary from 0 to as many as 5 (Table
5.1.1a). Typically one joint set cuts the rock mass into plates, two perpendicular sets cut
rock into column and three into blocks, and more sets cut rocks into mixed shapes of
blocks and wedges, as shown in Figure 5.1.1a.

The mechanical properties of the rock mass is obviously influenced by the presence of
joint sets and the number of joint sets. More joint sets provide more possibilities of
potential slide planes for rock wedges or blocks to slide and fall.

Figure 5.1.1a Rock masses showing one and three joint sets.

Table 5.1.1a ISRM suggested description of joint sets


I Massive, occasional random fractures
II One joint set
III One joint set plus random fractures
IV Two joint sets
V Two joint sets plus random fractures
VI Three joint sets
VII Three joint sets plus random fractures
VIII Four or more joint sets
IX Crushed rock, earth-like

Different from joints, rock fractures are considered as a non-systematic discontinuous


feature of rock masses. They are not in sets or parallel. They could be large in term of
numbers but their distribution is generally random. Rock mass quality is influenced by
the number of rock fractures and they are usually considered in the overall degree of
fracturing of a rock mass, in term of joint spacing and RQD, discussed in later sections.
Chapter 5 Properties of Rock Discontinuities 2

Persistence is the areal extent or size of a discontinuity, and can be crudely quantified by
observing the trace lengths of discontinuities on exposed surfaces. The persistence of
joint sets controls large scale sliding or 'down-stepping' failure of slope, dam foundation
and tunnel excavation. Figure 5.1.1b gives diagrams showing persistence of various
joint sets, while Table 5.1.1b presents the classification of persistence commonly adopted.

Figure 5.1.1b Sketches indicating persistence of various joint sets.

Table 5.1.1b ISRM classification of discontinuity persistence


Description Surface Trace Length (m)
Very low persistence <1
Low persistence 1–3
Medium persistence 3 – 10
High persistence 10 – 20
Very high persistence > 20

5.1.2 Joint Orientation: Joint Plane Orientation and Representation

Orientation of a discontinuity is described by its dip and dip direction or its dip and strike.
The orientation of major joint set relative to an engineering structure largely controls the
possibility of unstable conditions or excessive deformations developing. The mutual
orientation of discontinuities will determine the shape of the individual blocks and beds
comprising the rock mass.

Orientation of a plane is measured by the degree of inclination and the direction of facing
of the plane. It does not fix its position. Therefore, two parallel planes have the same
orientation. In rock mechanics and engineering geology, the orientation of a plane is
generally defined by dip angle (inclination), dip direction (facing) or strike (running), as
illustrated in Figure 5.1.2a.

Figure 5.1.2a Representation of joint plane orientation.

Dip or dip angle represents the degree of inclination. It is the acute angle between the
plane and the horizontal plane. It is also the acute angle between a line with maximum
dip in the inclined plane and its horizontal projection. Dip angle is generally expressed
by an acute angle between 0° and 90°.

Dip direction represents the facing direction. It is the bearing measured clockwise from
the north (0°) of the line with maximum dip in the inclined plane. Dip direction is
generally expressed by a direction angle of 0° to 360°.
Chapter 5 Properties of Rock Discontinuities 3

Strike is the alignment or run. It is the bearing of an imaginary horizontal line in the
inclined plane. Strike is generally expressed by a direction angle of 0° to 180°.

Dip direction and strike direction are always perpendicular. In rock mechanics, dip
direction/dip format is generally used, e.g., 210/35, or 030/35, where dip directions
always have 3 digitals. Sometime, when strike is used instead of dip direction, the
general direction of plane dip must be given, otherwise, it could means two possible
planes, e.g., dip/strike 120/35 would be either dip direction/dip 210/35, or 030/35.
Therefore correctly it should be presented as strike/dip 120/35SW which is the plane in
dip direction/dip 210/35, or 120/35NE which is the plane in dip direction/dip 030/35.

Normal to the plane is the imaginary line at right angle to the plane. Therefore the

trend of normal = dip direction of the plane ± 180,


orientation of the normal is given by,

plunge of normal = 90 – dip.

Orientation of a joint plane can be represented graphically using hemispherical projection


method. The projection method is to represent a 3D plane by a 2D presentation. The
most common projection is the low hemispherical equal angle projection. Use the
projection, joint orientation data can be assessed in 2D form.

Figure 5.1.2b Analysis of joint orientation data using projection method.

It is a powerful tool to analyse large number of joint data and examine the rock slope
stability, slide of rock block in underground excavation, stability of rock foundation on
jointed rock mass. The use of the hemispherical projection method is given in a later
section in this chapter.

5.1.3 Joint Spacing: Joint Spacing, Frequency, Block Size, and RQD

The degree of fracturing of a rock mass is controlled by the number of joint in a given
dimension. A rock mass contains more joints is also considered as more fractured. More
joints also mean that average spacing between joints is less. Several parameters can be
used to express the fracturing degree of a rock mass.

The spacing of adjacent joints largely controls the size of individual blocks of intact rock.
It controls the mode of failure. A close spacing gives low mass cohesion and circular or
even flow failure. It also influences the mass permeability.

Joint spacing for a particular pair of joint is the perpendicular distance between the two
joints. For a joint set, is usually expressed as the mean spacing of that joint set.
However, when the expose is limited, often the apparent spacing is measured. Figure
5.1.3a shows the relationship between spacing of individual joint set, apparent spacing
and average spacing. In the assessment of rock fracturing degree, the overall average
Chapter 5 Properties of Rock Discontinuities 4

spacing is considered. However, as illustrated in the figure, measurements of the overall


average joint spacing are different on different measuring faces.

Figure 5.1.3a Joint spacing, apparent spacing and true spacing.

ISRM recommends the use of the terms in Table 5.1.3a to describe joint spacing. The
description ranges from extremely close spacing to extremely wide spacing.

Table 5.1.3a Classification of discontinuity spacing


Description Joint Spacing (m)
Extremely close spacing < 0.02
Very close spacing 0.02 – 0.06
Close spacing 0.06 – 0.2
Moderate spacing 0.2 – 0.6
Wide spacing 0.6 – 2
Very wide spacing 2–6
Extremely wide spacing >6

Joint frequency (λ), is defined as number of joint per metre length. It is therefore simply
the inverse of joint spacing (sj), i.e.,

λ = 1 / sj

Another measure of fracturing degree is the Rock Quality Designation (RQD). Is is


defined as the percentage of rock cores that have length equal or greater than 100 mm
over the total drill length (Figure 5.1.3b).

ΣLength of cores >100 mm


RQD = × 100%
Total length of drilling

Figure 5.1.3b Example of measuring RQD from core logging.

Although RQD was initially proposed as an attempt to describe rock quality, in reality, it
only describes fracturing degree, by in fact considering the spacing of joints. Therefore,
statistically, RQD can be correlated to joint spacing or joint frequency the following
equation:
–0.1λ
RQD = 100 e (0.1λ +1)

For values of λ in the range 6 to 16/m, the above equation can be approximated by,
Chapter 5 Properties of Rock Discontinuities 5

RQD = 110.4 – 3.68λ

Joint space also defines the size of rock blocks in a rock mass. When a rock mass
contains more joints numbers, the joints have lower average spacing and smaller block
size. Block size can be classified by the volumetric joint count, Jv, defined as number of
joint per m3 volume of rock mass, as presented in Table 5.1.3b.

Table 5.1.3b ISRM suggested block size designations


Designation Volumetric Joint Count, joints/m3
Very large blocks <1
Large blocks 1–3
Medium-sized blocks 3 – 10
Small blocks 10 – 30
Very small blocks > 30
Crushed rock > 60

RQD can be related approximately to Jv by:

RQD = 115 – 3.3 Jv, for Jv between 4.5 and 30.

For Jv < 4.5, RQD is taken as 100%, and for Jv > 30, RQD is 0%.

5.1.4 Joint Surface and Opening: Roughness, Matching, Aperture and Filling

A joint is an interface face of two contacting surfaces. The surfaces can be smooth or
rough; they can be in good contact and matched, or they can be poorly contacted and
mismatched. The condition of contact also governs the aperture of the interface. The
interface can also be filled with intrusive or weathered materials.

Joint surface roughness is a measure of the inherent surface unevenness and waviness of
the discontinuity relative to its mean plane. The roughness is characterised by large
scale waviness and small scale unevenness of a discontinuity. It is the principal
governing factor the direction of shear displacement and shear strength, and in turn, the
stability of potentially sliding blocks.

Roughness can be distinguished between small scale surface irregularity or unevenness


and large scale undulation or waviness of the discontinuity surface, as illustrated in Figure
5.1.4a.

Figure 5.1.4a Definition of joint roughness at different scale.


Chapter 5 Properties of Rock Discontinuities 6

A classification of discontinuity roughness has been suggested by ISRM, and is


reproduced in Figure 5.1.4b. It describes the roughness first in metre scale (step,
undulating, and planar) and then in centimetre scale (rough, smooth, and slickensided).
The classification is useful to describe the joint surface but does not give any quantitative
measure.

Figure 5.1.4b Typical joint surface profile and suggested descriptions and
corresponding joint roughness coefficient (JRC) at different scales.

Another commonly used roughness classification is proposed by Barton, termed as Joint


Roughness Coefficient (JRC). JRC number is 0 for the smooth flat surface and 20 for
the very rough surface. The proposed JRC is reproduced in Figure 5.1.4b. Joint
roughness is affected by geometrical scale. In the JRC classification, the value of JRC
decreases with increasing size.

It should be noted that in realty, profiles of joint surfaces are 3D features (Figure 5.1.4c).
The above descriptions are 2D based. It is therefore suggested to take several linear
profiles of a surface for the description and JRC indexing.

Figure 5.1.4c 3D presentation of joint surface.

Joint surface is a rough profile that can be described by statistic method and fractal.
(A section on fractal describing surface profile.)

Fractal method is applicable not only in 2D (linear profile), but also in 3D (surface plane
profile), as shown in Figure 5.1.4d. It is a very powerful tool to quantify the surface
profile.
(More)

Figure 5.1.4d 3D joint surface profiles and fractal numbers.

However, a joint is an interface of two surfaces. The properties of a joint are therefore
controlled by the relative positioning of the two surfaces, in addition to the profiles of
both surfaces. For example, joints in fully contacted and interlocked positions has little
possibility of movement and is also difficult to shear, as compared to the same rough
joints in point contact where movement can easily occur. Often, joints are differentiated
as matched and mismatched (Figure 5.1.4e). A Joint Matching Coefficient (JMC) has
been suggested by considering the contact percentage of two surfaces, as shown in Figure
5.1.4f. JMC various from 0, representing completely mismatched with a few contact
points only in the joint interface, to 1, representing completely matched with fully in
contact of the joint.
Chapter 5 Properties of Rock Discontinuities 7

Figure 5.1.4e Matched and mismatched joint surface.

Figure 5.1.4f Scheme of Joint Matching Coefficient (JMC) for rock joints.

In a natural joint, it is very seldom that the two surfaces are in complete contact. There
usually exists a gap or an opening between the two surfaces. The perpendicular distance
separating the adjacent rock walls is termed as aperture. Descriptions of aperture are
suggested in Table 5.1.4a. Joint opening is either filled with air and water (open joint)
or with infill materials (filled joint), as illustrated in Figure 5.1.4g. Open or filled joints
with large apertures have low shear strength. Open aperture also associates with high
permeability and storage capacity.

Figure 5.1.4g Joint aperture and joint with filling.

Table 5.1.4a Classification of discontinuity aperture


Aperture Description
< 0.1 mm Very tight
0.1 ~ 0.25 mm Tight "Closed feature"
0.25 ~ 0.5 mm Partly open
0.5 ~ 2.5 mm Open
"Gapped feature"
2.5 ~ 10 mm Widely open
1 ~ 10 cm Very widely open
10 ~ 100 cm Extremely widely open "Open feature"
>1m Cavernous

Aperture can be separated by mechanical aperture or real aperture and equivalent


hydraulic aperture or conducting aperture. The later is particularly important when
permeability is concerned.

Filling is material in the rock discontinuities. The material separating the adjacent rock
walls of discontinuities. The wide range of physical behaviour depends on the
properties of the filling material. In general, filling affects the shear strength,
deformability and permeability of the discontinuities.

5.1.5 Correlation between Various Geometrical Properties

Figure 5.1.5a is an illustration of all the important geometrical properties of rock joints
and fractures. As all the features in a rock mass have undergone the same geological
processes, some of the geometrical features has certain degree of correlation.
Chapter 5 Properties of Rock Discontinuities 8

Figure 5.1.5a Illustration of various geometrical characteristics of rock joints and


fractures.

(Discussions on correlations between: joint set number and joint spacing/RQD, JRC and
aperture, etc)

5.2 Mechanical and Hydraulic Properties of Rock Joints and Fractures

5.2.1 Normal Stiffness and Displacement

Normal deformation characteristics and normal stiffness of rock joints are important
parameters for analysis and design. As discussed in an earlier chapter, a joint represents
a discontinuity of stress and displacement. A natural joint always has opening aperture
of less than 1 mm to a few mm. With increasing normal stresses, the opening closes,
and contact areas of the joint surfaces increase. Therefore as shown in Figure 5.2.1a, the
normal stress – normal displacement curve can be highly non-linear. The normal
stiffness, slope of the curve, is therefore not a constant.

Figure 5.2.1a Normal stress - normal displacement relation of joints in a granite

There are several mathematical models describing the normal stress – displacement

a hyperbolic relation between normal stress, σn, and normal displacement, dn,
relationship. In developing a joint element finite element model, Goodman (1976) used

σn – σni
σni
dn t
=A( )
dmax – dn

where dmax is the maximum possible closure, σni = a seating pressure defining the initial
normal stress conditions for measuring normal displacement, and A and t are
experimentally determined constants.

Based on a great number of laboratory experiments on matched rock fractures in dolorite,


limestone, siltstone and sandstone, Bandis et al. (1983) proposed a hyperbolic function to
express the normal effective stress-closure relation of a matched fracture. Assuming
positive signs for compression and fracture closure and negative signs for tension and
fracture opening, the normal effective stress-closure relation is,

σn =
kni dn
1 – (dn/dmax)
Chapter 5 Properties of Rock Discontinuities 9

or

σn
dn =
kni + (σn/dmax)

where σn is the normal effective stress, dn is the fracture closure, dmax is the maximum
allowable closure, kni is the normal stiffness of the fracture at initial stress. When
normal stress becomes infinite, fracture closure approaches the maximum allowable
fracture closure, and simultaneously, normal fracture specific stiffness becomes infinite.
The fracture becomes a welded interface. On the other hand, when normal stress is zero,
fracture closure becomes zero, and the corresponding normal fracture specific stiffness is
named as initial normal fracture specific stiffness. The initial normal stiffness (kni) and
maximum allowable closure (dmax) can be determined from regular static fracture
deformation tests or fracture properties, i.e., fracture wall compressive strength (JCS),
fracture roughness coefficient (JRC) and average aperture thickness (ai) at initial seating
normal stress, as described by Barton et al. (1985). The model is commonly known as
the BB (Barton-Bandis) model.

The above hyperbolic BB model of the fracture normal behaviour is commonly used in
rock mechanics and engineering. Under cyclic loading/unloading condition, the BB model
describes that the initial load and unload cycles may cause a hysteresis between them.
Successive load/unload cycles can continue to stiffen the fractures, and the BB model
eventually tends to a hyperbolic elastic model without the hysteresis between the load and
unload cycles.

On the other hand, in the laboratory experiments on mismatched rock fractures, Bandis et
al. (1983) also found that the mismatched rock fractures exhibit much reduced normal
stiffness, compared to the matched fractures. A semi-log function was used to fit the
normal stress-closure curves, as expressed in the following:

log σn = p + q dn

where σn is normal effective stress, dn is the fracture closure, p and q are material
constants.

Logarithmic functions have also been used by others to describe the normal behaviour of
rock fractures. For example, Zhao and Brown (1992) found that the normal stress -
normal displacement could be fitted by a function below,

dmax – dn
= 1 – A ln(σn/σni)
dmax – dni

where dni = displacement at a reference normal stress σni, usually equal to the seating
pressure, and A is constant varies from 0.16 to 0.21.
Chapter 5 Properties of Rock Discontinuities 10

The curve shown in Figure 5.2.1a indicates that at high normal stress, when the joint is
highly closed, the normal stiffness approaches that corresponding to the elastic modulus
of the rock material. When the joint is completely closed, there is no further closure of
the joint, the displacement is therefore only by the elastic deformation of the rock
material.

5.2.2 Shear Strength of Rock Joints and Fractures

Shear behaviour of rock joints is perhaps one of most important feature in civil
engineering rock mechanics. Conditions for sliding of rock blocks along existing joints
and faults at slope or excavation opening are governed by the shear strengths developed
on the sliding rock discontinuities. As seen in Figure 5.2.2a, in slope, shear is subjected
to a constant normal load generated by the weight of the blocks; while in tunnel, shear is
subjected to constant stiffness due to the constraints of lateral displacement.

Figure 5.2.2a Controlled normal load (a, c) and controlled normal displacement (b, d)
shearing modes and tests.

The shear properties are usually determined by direct shear test shown in Figure 5.2.2a.
Detailed description of test preparation and methodology is given in a later section.

surfaces gives the relationship between the friction angle φ, the normal force (N) and
As shown early in chapter on mechanics, sliding between two smooth horizontal contact

shear force (Fs), as Fs = N tanφ.

It is therefore not surprised that shear tests carried out on smooth, clean fracture surfaces
at controlled normal load condition generally give shear strength (s) - effective normal
stress (σn) curve (Figure 5.2.2b) and it follows the simple Coulomb law:

τ = σn tanφ

where φ is the effective angle of friction of the fracture surfaces.


Figure 5.2.2b, φ = 35°, a typical value for quartz-rich rocks.
For the case shown in

Figure 5.2.2b Shearing of smooth quartzite surfaces under various conditions.

Naturally occurring discontinuity surfaces are far from being smooth. Figure 5.2.2c is
typical of the results obtained for clean, rough fractures. As observed in the tests, shear
stress quickly mobilised and reaches a peak. When shearing is progressed, the shear
strength stablised to a residual level. The peak is usually term as the peak shear strength
and the residual is the residual shear strength. For rough joints, peak shears strength is
significantly higher than the residual strength.
Chapter 5 Properties of Rock Discontinuities 11

Figure 5.2.2c Results of a direct shear test on a clean rough rock joint.

Observations of shear test results show that residual strength follows the linear friction
law, i.e.,

τr = σn tan φr

On the other hand, peak shear strength does not follow the linear fiction law. The peak
strength for rough joints does not linearly proportional to normal stress. The gradient of
the peak shear strength – normal stress decreases with increasing normal stress.

As shown early in Chapter 3, for idealized rough fracture models by Patton (1966) shown
in Figure 5.2.2d, it is similar as sliding between two contact surface at an inclination.
Therefore, at low normal stress and at relatively short shear distance, shear strength is also
influenced by the inclination angle,

τ = σn tan(φ+i)

It was found that when the normal stress is increased above a critical value, shear stress
can eventually be developed so high that it causes shear failure through the asperities.
When such shearing through asperity occurs, the shear strength is somehow related to the
shear strength of the materials of the asperities. Comparing to rock joint, rock materials
have higher cohesion and internal friction angle of generally around 30°.

Figure 5.2.2d Idealized surface roughness models and bilinear peak strength envelope.

Therefore, shear strength for a rough fracture could exhibit two features, a lower portion
representing shearing by climbing the asperity angle, and an upper portion representing

5.2.2d, and is expressed by the equations below. In the equation, σn’ is the critical
shearing off the asperities. This leads to a bilinear shear strength model shown in Figure

normal stress when shearing of asperity is assumed to start.

σn tan (φ+i) for σn ≤ σn’


τ=
c + σn tan φ for σn ≥ σn’
{

However, in reality, there is not clear boundary between shearing by climbing the asperity
angle and shearing off the asperities. With increasing normal stress, asperity shearing
off increases progressively. Therefore, the actual shear stress – normal stress relation is
represented by a curve, as shown in Figure 5.2.2c.
Chapter 5 Properties of Rock Discontinuities 12

Based on extensive test results and noticing the progressive damage of asperities, Barton
(1973) proposed that the peak shear strengths of joints could be represented by the
empirical relation below,

τ = σn tan [JRC log10 JCS


) + φr]
( σn

where σn = effective normal stress, JRC = joint roughness coefficient on a scale of 1 for

φr = drained residual friction angle.


the smoothest to 20 for the roughest surfaces, JCS = joint wall compressive strength, and

(Discussion on dilation and dilation angle.)

5.2.3 Other Factors Affecting Joint Shear Behaviour

Roughness effect can cause shear strength to be a directional property. Figure 5.2.3a
illustrates a case in which rough discontinuity surfaces were prepared in slate specimens.
Directional effects are not just in foliated rocks, but rather universal. As discussed in the
geometrical properties, surface profile is a 3D feature while shearing is a directional
activity. Surface profile along a particular direction would be different along another
direction and hence gives different shear strength.

Figure 5.2.3a Effect of shearing direction on the shear strength of a joint in a slate.

The natural discontinuities normally suffered weathering and alteration, which in term,
also change the degree of matching of the discontinuity surfaces. It was found that the
mismatched discontinuities generally have much lower shear strength than matched
(interlocked) ones (Figure 5.2.3b).

Figure 5.2.3b Shear strength of matched and mismatched fractures in a granite.

When a joint is wet, it has generally a lower friction angle than a dry joint. The shear
strength of a wet joint is calculated use the wet friction angle. If the joint is subjected to
groundwater pressure, the normal stress in the shear strength equation is the effective
normal stress, i.e., total stress – water pressure.

both scale dependent and stress dependent. As σn increases, the term log10(JCS/σn)
The JRC-JCS shear strength equation shows that the shear strength of a rough joint is

decreases, and so the net apparent friction angle decreases. As the scale increases, the
steeper asperities shear off and the inclination of the controlling roughness decreases.
Similarly, the asperity failure component of roughness decreases with increasing scale
Chapter 5 Properties of Rock Discontinuities 13

because the material compressive strength, JCS, decreases with increasing size, as
illustrated in Figure 5.2.3c.

Figure 5.2.3c Influence of scale on the three components of discontinuity shear


strength.

5.2.4 Flow and Permeability of Rock Joints

From the early chapter on mechanics, it showed that flow in parallel plates is governed by
the cubic flow law. The parallel plates theory is applicable to flow in rock joints.
Therefore, flow and permeability of a rock joint are given as,
3

12 ν
w i g de
Q= (5.2.4a)

12 ν
g de
k= (5.2.4b)

where g = acceleration due to gravity, ν = kinematic viscosity of the fluid, w = width of


the joint, and d = aperture of smooth plates or equivalent hydraulic aperture of the rough
joint.

The parallel plates theory is assumed for smooth plates and laminar flow. When it is
applied to actual rock joints with rough surfaces, which are far from smooth, the equation
does not truly represent the real case. The original equation therefore, does not account
for the deviations from the ideal conditions due to the joint surface geometry and other
effects. Somehow, modification has to be introduced to reflect the effects of joint
roughness and flow path. Therefore, in the above equation, instead of the aperture of
smooth plates, in natural rock joints, equivalent hydraulic aperture is used. The equivalent
hydraulic aperture of a rock joint (de)is estimated from,

de = f d (5.2.4b)

where d is the actual aperture of the rock joint, and f is a factor that accounts for

and f ≤ 1.
deviations from the ideal conditions that are assumed in the parallel smooth plate theory,

It is found that for a given joint, f is a constant at different apertures, without change of
joint surface profile (Witherspoon et al 1980). It is also noted that f value is generally
lower when the joint surfaces are rougher. This means that rougher joints deviate more
from smooth parallel plates and hence require higher corrections.

5.3 Correlations between Geometrical, Mechanical and Hydraulic Properties


Chapter 5 Properties of Rock Discontinuities 14

5.3.1 Joint Surface Profile and Normal Stiffness

It was observed that closure under load was more complete in smooth joints than in rough
joints. Conversely, rough joints in strong rocks close least under normal stress. The initial
normal stiffness and maximum closure were dependent on roughness (JRC) and wall
strength (JCS).

The effect of joint surface mismatch was noticed. Earlier experiments performed by
Bandis (1980) suggested that when mismatch occurs the number of contact points may
reduce, although the individual areas of contacting asperities may become larger.

5.3.2 Joint Surface Profile and Shear Strength

The JRC-JCS joint shear strength criterion has already highlighted the relationship
between joint roughness and strength. It is evident that rougher joint surface leads to
higher shear strength.

(Discussion on correlation between fractal and shear strength.)

5.3.3 Joint Surface Profile and Permeability

Many studies have been conducted on strength, deformation and conductivity coupling of
rock joints in an attempt to relate these to the joint surface roughness. A relationship
between equivalent hydraulic aperture and real joint aperture based on the Joint
Roughness Coefficient (JRC) was proposed by Barton and Choubey [1977]:
2.5
JRC
de = 2 (5.2.5b)
(d/de)

where de is the equivalent hydraulic aperture and d is the real aperture of a joint.

5.3.4 Joint Closure and Permeability

The permeability and hydraulic aperture of rock joints changes with effective normal
stress. As shown in Figure 5.3.4a, joint permeability reduces asymptotically and
approaches to zero with increasing effective normal stress.

Figure 5.3.4a Changes of permeability with effective normal stress of rock joints in a
granite.
Chapter 5 Properties of Rock Discontinuities 15

A hydraulic model describing the hydraulic behaviour of discontinuities was proposed by


Walsh (1981) and modified by Zhao and Brown (1992). The model suggested a
logarithmic relation between the joint permeability, kj and the effective normal stress,
(σn′),

σ′ 2
σ r′
kj
= [1 – B Ln ( n ) ] (5.3.4a)
kr

where kr = the rock joint permeability at a reference effective normal stress σr′, and B is a
parameter dependent on surface properties of the joint.

5.3.5 Joint Shear, Aperture and Permeability

For an originally matched and closed joint, shear will start to general separation of the
joint surface and creating larger aperture and high permeability, as illustrated in Figure
5.3.5a. As seen from the figure, when shear occurs, dilation occurs due the climbing
effects. The climbing effects may be less obvious if the joint is under high normal
stress. In this case, the asperities would be crashed and crashed particles may be filled in
the joint. This may still result in increasing of permeability but not as significant as in the
previous case.

Figure 5.3.5a Change of aperture with shear displacement of a matched joint.

For a non-matched joint, the situation may be quite different. Depending on the original
situation, the aperture could be reduced if shearing of the joint causes close up of the
joint, or vice versa.

5.4 Behaviour of Joints under Cyclic and Dynamic Loading

5.4.1 Joint Surface Damage under Cyclic Loading

5.4.2 Joint Behaviour under Dynamic Loads

5.4.3 Factors affect Rate Dependent Characteristics of Joints

5.5 Effects of Joints on Transient Stress Wave Propagation

5.5.1 General Concept of Dynamic Stress and Transient Waves


Chapter 5 Properties of Rock Discontinuities 16

5.5.2 Effects of Single Joint on Wave Transmission

5.5.3 Effects of Joint Set on Wave Transmission

5.6 Characteristics of Rock Faults and Folds

5.6.1 Single Fault

Single fault should be characterised similarly as joint, including orientation, persistence,


surface roughness, aperture and filling. Persistence or length of the fault is particularly
important in order to appreciate the impact and influence of the fault.

Another aspect of importance is groundwater flow in the fault. Faults are usually of
great length; they generally are better connected than most of the joints, and hence create
a water flow channel.

5.6.2 Fault Zone of Extended Thickness

In addition to the characteristics of planer fault, thickness of a fault zone has important
influence on the overall properties. Together with the thickness, the materials within the
fault zone should be properly described and understood. The materials can vary from
crushed to completely decomposed rocks. The properties of those materials need to be
tested and determined in order to estimate the strength and deformation characteristics.

Similarly to single fault, fault zones also often become major groundwater flow channel.
Major faults sometimes are associated with and connected to surface geographic
depression and water body.

5.6.3 Bedding Planes and Rock Formation Interfaces

Bedding planes of sedimentary rocks without being folded are planner. Important
characteristics need to be described are the orientation and interface types.

In most cases, conformable or unconformable bedding planes are cemented and do not
represent a separation with an opening. Unconformable bedding planes may be
represented by a mixed interface in which materials of both rocks of each side are mixed
and hence dose not show a clear line separating the two rocks.

Non-conformable interfaces are the interfaces between sedimentary rocks with non-
sedimentary (igneous and metamorphic) rocks. They may not be planner, and may be
Chapter 5 Properties of Rock Discontinuities 17

represented by mixed interfaces containing fragments of rocks on both sides, or may be


represented by localised contact metamorphism caused by intrusion.

Dykes and sills are localised intrusions of igneous materials into existing rocks. The
interfaces between dykes/sills with the existing rocks are represented by contact
metamorphism.

Interfaces between two non-sedimentary rocks are usually well welded, by intrusion or by
metamorphism. The interfaces therefore only represent a discontinuity of materials but
not necessarily a weak zone or failure plane.

The condition of rocks, particularly carbonate sedimentary rocks (limestone and dolomite)
close to the interface needs to be carefully examines. For example, at an interface
between porous sandstone and limestone with active groundwater flow, limestone may be
weathered and showing well developed cavities.

5.6.4 Intensively Folded Thin Layers

Sedimentary layers of relative thin thickness and intensively folded often represent a zone
of fractured and weak rock. Description of discontinuities is not easy. However,
general descriptions should include the layer thickness, materials in the layers, degree and
type of folding, and groundwater condition.

In the Chapter dealing with rock mass, such zones will be discussed in term of rock mass
classification.

5.7 Field and Laboratory Characterisation of Rock Joints

5.7.1 Overview on Field and Laboratory Methods

Characterisation of rock discontinuities are done by three means, most convenient and
best mean is by mapping at outcrops. Therefore outcrop mapping should always be the
first choice of exposure of rock face is available. Rock cores from boreholes provides
many useful information on rock discontinuities, and core logging remains an important
exercise of rock discontinuity characterisation. In addition to core logging, further
information can often be supplemented by log the borehole. Geophysical borehole
logging becomes increasingly useful in rock discontinuity and rock mass characterisation.

Table 5.7.1a provides an overview on the applicability of various methods to measure


rock discontinuities from outcrop mapping and core logging.

Table 5.7.1a Measurement of discontinuity geometrical features


Feature Measurement Method Outcrop Core Borehole
Mapping Logging Logging
Chapter 5 Properties of Rock Discontinuities 18

Discontinuities type Visual good good medium


Orientation Compass-clinometer good medium good
Spacing Measuring tape good good medium
Persistence Measuring tape good poor poor
Roughness Profile gauge good medium poor
Wall strength Schmidt hammer good medium poor
Aperture Scale or feeler gauge good poor poor
Filling Visual good poor poor
Seepage Timed observation good poor good
Number of joint sets Hemispherical projection good medium poor
Block size 3-D fracture frequency good poor poor

5.7.2 Identification of Joint Sets

Measurements on joint set number are usually done by observation and orientation
measurements at outcrops.

Descriptions of joint sets are suggested by ISRM, as reproduced in Table 5.7.2a.

Table 5.7.2a ISRM suggested description of joint sets


I Massive, occasional random fractures
II One joint set
III One joint set plus random fractures
IV Two joint sets
V Two joint sets plus random fractures
VI Three joint sets
VII Three joint sets plus random fractures
VIII Four or more joint sets
IX Crushed rock, earth-like

It is not easy to measure joint set number by logging the rock cores. Often dominating
joint sets or joint sets most perpendicular to drilling can be identified. Joints parallel and
sub-parallel to drilling are not well represented in core and hence not easily notified.

5.7.3 Measurement of Joint Orientation

(a) By Outcrop Mapping

The most convenient way to measure joint orientation is from accessible outcrops or
exposed faces of slope cuts or underground excavation. The measurements can be made
by a geological compass, which gives readings of dip direction (bearing) and dip angle
(inclination), as shown in Figure 5.7.3a.
Chapter 5 Properties of Rock Discontinuities 19

Orientation of a joint plane daylighted on exposed surfaces may be obtained by surveying


methods from an inaccessible outcrop. The measurement may give orientations of the
daylighted lines. Orientation of the joint plane can be calculated from the orientations of
the daylighted traces of the same joint plane, as shown in Figure 5.7.3b.

Assume the orientations of the two trace lines are α1, β1, and α2, β2 (plunge and trend),
from 3D geometry, the orientation of the joint plane (dip angle α, dip direction β) is given
by the equation below,

tan α1
= cos (|β – β1|)
tan α

and

tan α2
= cos (|β – β2|)
tan α

By combining the above two equations, we have,

tan α1 cos (|β – β1|)


tan α2 cos (|β – β2|)
=

With given α1, β1, and α2, β2, dip direction of the plane β can be calculated by the above
equation. Dip angle α can be calculated by substitute β to one of the earlier equations.

The determination of plane orientation from the two daylighted lines can also be done by
projection method, which will be presented in a later section in this Chapter.

The dip angle shown by the trace of the daylighted joint plane is called apparent dip.
Apparent dip is always smaller then the true dip, as the true dip is defined as the
maximum dip angle of the plane.

(b) By Core and Borehole Logging

Joint are intersected by borehole drilling and hence can be seen from the cores obtained
from coring. Boreholes mostly are drilled vertically. Therefore, dip angle of joints and
fractured can be easily estimated, as the angle between the joint plane (when core is
placed vertically) and the horizontal. However, drilling is by rotational coring and
usually the bearing of cores is not fixed. Therefore, the dip direction cannot be
determined, in normal drilling.

Dip direction determination is possible if core orientation is known. Core orientation is


possible in reasonably good quality rock, where joints are reasonable close and matched.
mark, indicating, say, north, is printed on the core before drilling and when the cores are
taken out and reconnected, the whole core samples can be reoriented and dip directions of
all the joints and fractures can be determined, as illustrated in Figure 5.7.3c.
Chapter 5 Properties of Rock Discontinuities 20

In inclined and horizontal drilling, core orientation can be done within a drilling system.
The core barrel can have a steel ball which sit at the lowest position, i.e., lower side of the
core. The steel ball is locked in the core barrel and kept therefore the in the same
orientation as the cores. When the cores are taken out from the borehole, cores can be
reoriented with the aid of the steel ball, as shown in Figure 5.7.3d.

Orientation can also be determined by log the borehole, for example, by impression
packer or acoustic imaging. Those methods are aimed at obtaining the images of the
borehole walls. The images can be reconstructed to produce the joint plane cutting
through the borehole. With know orientation of the image, the orientation of the joint
can be easily determined, as shown in Figure 5.7.3e.

5.7.4 Measurement of Joint Spacing and RQD

(a) By Outcrop Mapping

At an outcrop where rock is exposed, a scanline, say, horizonally along a straight outcrop
surface is planed. Along the scanline, using a measuring tape, spacing of joint
daylighted on the outcrop can be measured. Measurements can be done in three ways:
(a) measuring the total amount of joint numbers with the scanline length, to calculate the
joint frequency; (b) measuring all the individual spacing between all the joints, to
calculate average spacing of all the joints: (c) measuring spacing of joints of individual
joint sets, to calculate joing spacing for different joint sets; and (d) measuring all the
spacing longer than 10 cm, to calculate RQD. Various measurements are illustrated in
Figure 5.7.4a.

It should be noted that the measurements on the outcrop surface give the apparent spacing
of joints. The measurements are also directional, i.e., if the scanline is in different
direction, say vertical, the measurements will be different.

(b) By Core and Borehole Logging

Measuring RQD is almost a standard practice during core logging. It is usually


measured for each core run (generally 1 – 3 m), or for the length of cores in a core box
(generally 1 – 1.5 m). By placing a measuring tape along one side of the core length,
rock cores have a length longer than 10 cm are noted and summed, dividing to the drilling
length, giving the RQD. Alternatively, the total number of fractures can be counted to
calculate the joint frequency. The measurements are illustrated in Figure 5.7.4b.

In core logging for RQD or frequency, the length to be divided is the total drilling length,
not the core length. In competent rock and with good drilling practice, the core length
can be the same as drilling length. Sometimes, rock cores are not fully recovered from
drilling, and then the core length is shorter than the drilling length. The ratio of
recovered core length to the drilling length is termed as core recovery. When coring
Chapter 5 Properties of Rock Discontinuities 21

through a highly fractured rock mass or a faulted zone, core recovery could be low due to
loss of loss materials in the fractured and faulted zones.

5.7.5 Joint Surface Profile Measurements

(a) Measurement of Large Scale Waviness at Site

Large scale waviness of a joint at site can be obtained by placing a long ruler over the
joint surface and then to measurement at a fixed interval the gap between the ruler and the
profile surface, as indicated by illustration in Figure 5.7.5a.

(b) Measurement of Roughness at Small Scale

Roughness measurements are usually done by a profile gauge shown in Figure 5.7.5b.
More precise measurement can be obtained by using a laser device, as shown in Figure
5.7.5c. A simple profile gauge provides a profile along a scanline and each profile is
then compared with a typical profile to give the roughness description or the roughness
number. Alternatively, fractal number can be computed.

With a laser profile capable to move along x and y directions, a series linear profiles can
be scanned to provide a 3D profile plane. With the 2D profile or 3D profile, toughness
can be described, or fractal numbers be calculated.

5.7.6 Description of Joint Surface and Filling

(a) Weathering and Alteration

Weathering and alternation is usually visible at outcrops or from the cores. When the
joint surface is weathered, it often shows the change of colour and appearance. Often,
weathered products, such as grain particles may also remain inside the joint. Detailed
description is necessary. Table 5.7.6a gives the suggested description by ISRM.

Table 5.7.6a ISRM suggested descriptive terms for joint surface alteration

Term Description
Fresh No visible sign of weathering of rock material at joint wall.
Discoloured Colour of the original fresh rock material is changed. The
degree of change from the original colour should be indicated.
If the colour change is confined to particular minerals this
should be mentioned.
Decomposed Rock is weathered to the condition of a soil in which the
original materials fabric is still intact, but some or all of the
mineral grains are decomposed.
Chapter 5 Properties of Rock Discontinuities 22

Disintegrated Rock is weathered to the condition of a soil in which the


original materials fabric is still intact. The rock is friable, but
the mineral grains are not decomposed.

(b) Filling in Joint

Joint can be clean or filled with weathered products and deposits, ranging from sandy
particles to swelling clays. Descriptions of filling materials need be given in details, in
term types of the materials, thickness, and particle sizes. If swelling clays are found,
swelling characteristics should be described.

(c) Estimating Joint Wall Strength

Joint wall strength is also an indicating of weathering and alteration of joint wall. When
the joint is weathered, the strength of the rock at joint wall reduces significantly. As we
discussed earlier, this affects greatly the shear strength of the joint.

Joint wall strength can be estimated by a Schmidt hammer. With the Schmidt hammer
number, uniaxial compressive strength can be estimated.

5.7.7 Estimation of Joint Aperture and Contact Areas

(a) By Outcrop Mapping and Core Logging

At outcrop mapping, joint aperture can only be roughly estimated, through direct
observation of joint exposed at outcrop, according to the ISRM suggested description
represented in Table 5.1.4a. The actual measurement is rather difficult, if not
impossible.

(b) By Laboratory Measurements

Specific methods have been developed in the laboratory to measure the aperture and
contact area of rock joints. The most common method is by impress trace. Materials are
injected into the joint and are allowed to set. When the joint is opened, the hardened
injected material gives the impression of the joint, including gaps and contacts. Contact
points and areas as well as aperture can then be estimated.

5.7.8 Permeability Measurements of Rock Joints

(a) In Situ Tests


Chapter 5 Properties of Rock Discontinuities 23

In situ permeability tests usually are done in boreholes for a section of rock mass, and
they will be described in details in the next Chapter. For measuring permeability of
individual joint, tests can be done in a borehole with packers.

From core or borehole logging, the joint to be tested should be selected. The joint
should be able to be isolated by a pair and packer and between the packers, there should
be only that joint within the tested section. A pair of packers are lowered down into the
borehole to the positions, to include the joint between the packers. The packers are
inflated to seal the section. Permeability tests are conducted by injecting high pressure
water within the section sealed by the packers. The test is often referred as borehole
packer test, and is illustrated in Figure5.7.8a. Permeability (often expressed as
transmissivity) can be calculated from flow characteristics, flow transmitting rate and
flow pressure.

(b) Laboratory Tests

Permeability tests on joint in laboratory can be set up using a system similar to Darcy’s
experiment. In addition, normal stress may be applied to the joint to determine the flow
rate and permeability at various stress conditions. A typical set-up using a triaxial cell is
shown in Figure 5.7.8b.

Permeability can be calculated from the flow rate measurements, hydraulic gradient and
specimen geometry, when the water flow is steady state laminar flow in the joint. Using
the parallel plates theory, equivalent hydraulic aperture can be estimated.

Change of pressure in the cell causes change of normal stress acting on the joint, and
leads to change of joint aperture. Such change will also be reflected in the change of
permeability.

5.7.9 Normal Compression and Stiffness Measurement of Joints

Rock sample containing a joint is prepared. Ideally, the joint should be placed
horizontally, parallel to the loading plane. The specimen can be cut into circular cylinder
or rectangular block and cross section area is measured. The joint surface is carefully
protected from mechanical damage during cutting and preparation. The profiles of joint
surfaces are recorded using a profiling gauge. The specimen is loaded under a standard
compression machine with load measurement. LVDTs or dial gauges are placed near
and across the joint to measure the normal displacement of the section containing the
joints, as shown in Figure 5.7.9a.

Load and displacement measurement should be taken regularly. If the displacement are
measured a relative large section of the rock, the displacement of the rock material should
be subtracted from the total displacement to give the net displacement of the joint.
Chapter 5 Properties of Rock Discontinuities 24

Stress (load/cross-section area) and joint normal displacement are plotted to give the
stress-normal displacement behaviour of the joint. Normal stiffness at a specific stress
level is the gradient of the tangent to the stress-normal displacement curve at that stress,
as illustrated in Figure 5.7.9b. It should be noted that the stress-normal displacement
behaviour of a rough joint is a curve.

5.7.10 Direct Shear Strength Test of Joints

Rock sample containing discontinuity is prepared and encapsulated in laboratory shear


box, with the discontinuity laid horizontally. The discontinuity is carefully protected
from mechanical damage during cutting and preparation. The sample is then mounted in
shear box using plaster, as shown in Figure 5.7.10a. The profile of discontinuity surface
are recorded using a profiling gauge. Area of the discontinuity is also measured. The
discontinuity is loaded under a constant normal load, and shear force is applied using a
mechanical gear-drive system (Figure 5.7.10b). Shear displacement, shear force and
normal displacement are recorded at a constant shear displacement interval (0.2-0.25
mm). The tests are continued until residual shear strength is obtained or about 10% of
the specimen length (Figure 5.7.10c).

Normal stress (σn), peak shear strength (τp) and residual shear strength (τr) are calculated
as normal load, peak shear force and residual shear force divided by the shear area.
Peak shear strength, normal stress and angle of friction (φ) can be adjusted to account for

displacement curve (h) as i = δn / δh


dilation. The angle of dilation (i) is estimated from normal displacement (n) - shear

Adjusted basic angle of friction (φ′) = ( φ − i ).


Adjusted normal stress (σn′) = ( σn cos i + τp sin i ) cos i
Adjusted peak shear strength (τp′) = ( τp cos i − σn sin i ) cos i

Reporting of results includes description of rock specimen and discontinuity, surface


roughness profile, shear stress - shear displacement and normal displacement - shear
displacement curves, peak shear strength, residual shear strength at each normal stress,
plots of peak shear strength and residual shear strength against normal stress .

5.8 Hemispherical Projection Method

5.8.1 Principle of Projection

5.8.2 Projection of Planes and Lines

5.8.3 Use of Projection for Geometrical Analysis


Chapter 5 Properties of Rock Discontinuities 25

5.8.4 Applications of Projection Methods


Chapter 5 Properties of Rock Discontinuities 26

5.7.8 Permeability Measurements of Rock Joints

(a) In Situ Tests

In situ permeability tests usually are done in boreholes for a section of rock mass, and
they will be described in details in the next Chapter. For measuring permeability of
individual joint, tests can be done in a borehole with packers.

From core or borehole logging, the joint to be tested should be selected. The joint
should be able to be isolated by a pair and packer and between the packers, there should
be only that joint within the tested section. A pair of packers are lowered down into the
borehole to the positions, to include the joint between the packers. The packers are
inflated to seal the section. Permeability tests are conducted by injecting high pressure
water within the section sealed by the packers. The test is often referred as borehole
packer test, and is illustrated in Figure5.7.8a. Permeability (often expressed as
transmissivity) can be calculated from flow characteristics, flow transmitting rate and
flow pressure.

The basic injection flow test procedures are outlined below:


(a) Open water feeding system valve and maintain constant pressure (PA), record the
elapsed time and total volume of consumed water every 0.5 minute, for the first 3
minute, then every minute, for about 10-15 minutes, until the pressure appears to
have stabilised.
(b) After pressure PA has stabilised for approximately 3 minutes, increase the water
pressure to pressure PB. Record the time and flow the same way as for PA, for about
10-15 minutes, until the pressure appears to have stabilised.
(c) After pressure PB has stabilised for approximately 3 minutes, increase the water
pressure to pressure PC. Repeating the same procedure by recording the time and
flow until pressure stabilised.
(d) Continue the tests for pressures PD and PE, following the same procedure.
5.1.1a

5.1.1a

1
N
St rike

Dip direct ion Vert ical plane


Dip angle
Measured on vertical plane: 55
N
Line of m axim um dip
Horizont al plane
Orient at ion:
Measured clockwise on horizontal plane: 220 Dip direct ion / Dip
220/ 55

5.1.2a

5.1.2b

2
Apparent spacing Apparent spacing Apparent spacing True spacing
On the plane in x direction in y direction

5.1.3a

5.1.3b

3
5.1.4a

5.1.4b

4
5.1.4c

5.1.4d

5
5.1.4e

5.1.4f

6
5.1.4g

5.1.5a

7
5.2.1a

(a) (b)

(c)

(d)

5.2.2a

8
5.2.2b

(a)
Shear Force

Shear Displacement

(b)
Shear Strength

Peak

Residual

Normal Stress
5.2.2c

9
N N
S S

N N
S S
i
i

φ
S S S

φ N φ+i N φ+i N
5.2.2d

5.2.3a

10
5.2.3b

5.2.3c

11
5.3.4a

5.3.5a

12
5.7.3a

Orientation of the
joint daylighted

Apparent spacing on Apparent spacing


the measuring surface in 3 directions

5.7.3b

13
RQD = (L1 + L2 + … + Ln) / L x 100%

λ = number of joints / length = n / L

Outcrop Face

1 2 Tape i n

L1 L2 L3 X L4 X X L5 Li X X Ln
L
<10 cm <10 cm <10 cm fault

5.7.4a

<10 cm <10 cm <10 cm core loss

L1 L2 L3 L4 L5 Li Ln
X X X X X
L

RQD = (L1 + L2 + … + Ln) / L x 100%


5.7.4b

14
Load measured
by load cell

Displacements
measured by
LVDTs

5.7.9a

15
Chapter 6 Rock Mass Properties and Classifications 1

CHAPTER 6
ROCK MASS PROPERTIES AND
CLASSIFICATIONS
Rock mass property is governed by the properties of intact rock materials and of the
discontinuities in the rock. The behaviour if rock mass is also influenced by the
conditions the rock mass is subjected to, primarily the in situ stress and groundwater.
The quality of a rock mass quality can be quantified by means of rock mass
classifications. This Chapter addresses rock mass properties and rock mass
classifications.

6.1 Rock Mass Properties and Quality

6.1.1 Properties Governing Rock Mass Behaviour

Rock mass is a matrix consisting of rock material and rock discontinuities. As discussed
early, rock discontinuity that distributed extensively in a rock mass is predominantly
joints. Faults, bedding planes and dyke intrusions are localised features and therefore
are dealt individually. Properties of rock mass therefore are governed by the parameters
of rock joints and rock material, as well as boundary conditions, as listed in Table 6.1.1a.

Table 6.1.1a Prime parameters governing rock mass property


Joint Parameters Material Parameters Boundary Conditions
Number of joint sets Compressive strength Groundwater pressure and
Orientation Modulus of elasticity flow
Spacing In situ stress
Aperture
Surface roughness
Weathering and alteration

The behaviour of rock changes from continuous elastic of intact rock materials to
discontinues running of highly fractured rock masses. The existence of rock joints and
other discontinuities plays important role in governing the behaviour and properties of the
rock mass, as illustrated in Figure 6.1.1a. Chapter 4 has covered the properties of intact
rock materials, and Chapter 5 has dealt with rocks contains 1 or 2 localised joints with
emphasis on the properties of joints. When a rock mass contains several joints, the rock
mass can be treated a jointed rock mass, and sometimes also termed a Hoek-Brown rock
mass, that can be described by the Hoek-Brown criterion (discussed later).

6.1.2 Classification by Rock Load Factor (Terzaghi 1946)

Based in extensive experiences in steel arch supported rail tunnels in the Alps, Terzaghi
(1946) classified rock mass by mean of Rock Load Factor. The rock mass is classified
Chapter 6 Rock Mass Properties and Classifications 2

into 9 classes from hard and intact rock to blocky, and to squeezing rock. The concept
used in this classification system is to estimate the rock load to be carried by the steel
arches installed to support a tunnel, as illustrated in Figure 6.1.2a. The classification is
presented by Table 6.1.2a.

Figure 6.1.2a Terzaghi’s rock load concept.

For obtaining the support pressure (p) from the rock load factor (Hp), Terzagh suggested
the equation below,

p = Hp γ H

where γ is the unit weight of the rock mass, H is the tunnel depth or thickness of the
overburden.

Attempts have been made to link Rock Load Factor classification to RQD. As suggested
by Deere (1970), Class I is corresponding to RQD 95-100%, Class II to RQD 90-99%,
Class III to RQD 85-95%, and Class IV to RQD 75-85%.

Singh and Goel (1999) gave the following comments to the Rock Load Factor
classification:
(a) It provides reasonable support pressure estimates for small tunnels with diameter up
to 6 metres.
(b) It gives over-estimates for large tunnels with diameter above 6 metres.
(c) The estimated support pressure has a wide range for squeezing and swelling rock
conditions for a meaningful application.

6.1.3 Classification by Active Span and Stand-Up Time (Stini 1950, Lauffer 1958)

The concept of active span and stand-up time is illustrated in Figure 6.1.3a and Figure
6.1.3b. Active span is in fact the largest dimension of the unsupported tunnel section.
Stand-up time is the length of time which an excavated opening with a given active span
can stand without any mean of support or reinforcement. Rock classes from A to G are
assigned according to the stand-up time for a given active span. Use of active span and
stand-up time will be further discussed in later sections.

Figure 6.1.3a Definition of active span.

Figure 6.1.3b Relationship between active span and stand-up time and rock mass
classes. Class A is very good and Class G is very poor.
Chapter 6 Rock Mass Properties and Classifications 3

Table 6.1.2a Rock class and rock load factor classification by Terzaghi for steel arch supported tunnels
Rock Load Factor Hp (feet)
Rock Class Definition Remark
(B and Ht in feet)
Hard and intact rock contains no joints and fractures. After
Light lining required only if spalling or
I. Hard and intact excavation the rock may have popping and spalling at excavated 0
popping occurs.
face.
Light support for protection against
II. Hard stratified Hard rock consists of thick strata and layers. Interface between strata
0 to 0.5 B spalling. Load may change between
and schistose is cemented. Popping and spalling at excavated face is common.
layers.
Massive rock contains widely spaced joints and fractures. Block size
III. Massive, Light support for protection against
is large. Joints are interlocked. Vertical walls do not require support. 0 to 0.25 B
moderately jointed spalling.
Spalling may occur.
Rock contains moderately spaced joints. Rock is not chemically
IV. Moderately
weathered and altered. Joints are not well interlocked and have small 0.25 B to 0.35 (B + Ht) No side pressure.
blocky and seamy
apertures. Vertical walls do not require support. Spalling may occur.
Rock is not chemically weathered, and contains closely spaced
V. Very blocky and
joints. Joints have large apertures and appear separated. Vertical (0.35 to 1.1) (B + Ht) Little or no side pressure.
seamy
walls need support.
VI. Completely Rock is not chemically weathered, and highly fractured with small Considerable side pressure. Softening
crushed but fragments. The fragments are loose and not interlocked. Excavation 1.1 (B + Ht) effects by water at tunnel base. Use
chemically intact face in this material needs considerable support. circular ribs or support rib lower end.
VII. Squeezing rock Rock slowly advances into the tunnel without perceptible increase in
(1.1 to 2.1) (B + Ht)
at moderate depth volume. Moderate depth is considered as 150 ~ 1000 m. Heavy side pressure. Invert struts
VIII. Squeezing rock Rock slowly advances into the tunnel without perceptible increase in required. Circular ribs recommended.
(2.1 to 4.5) (B + Ht)
at great depth volume. Great depth is considered as more than 1000 m.
Rock volume expands (and advances into the tunnel) due to swelling up to 250 feet, irrespective Circular ribs required. In extreme cases
IX. Swelling rock
of clay minerals in the rock at the presence of moisture. of B and Ht use yielding support.
Notes: The tunnel is assumed to be below groundwater table. For tunnel above water tunnel, Hp for Classes IV to VI reduces 50%.
The tunnel is assumed excavated by blasting. For tunnel boring machine and roadheader excavated tunnel, Hp for Classes II to VI reduces 20-25%.
Chapter 6 Rock Mass Properties and Classifications 4

6.1.4 Rock Quality Designation (RQD) (Deere 1964)

Rock quality designation (RQD) was introduced in 1960s, as an attempt to quantify rock
mass quality. Table 6.1.2a reproduces the proposed expression of rock mass quality
classification according to RQD.

As discussed earlier, RQD only represents the degree of fracturing of the rock mass. It
does not account for the strength of the rock or mechanical and other geometrical
properties of the joints. Therefore, RQD partially reflecting the rock mass quality.

Table 6.1.2a Rock mass quality classification according to RQD


RQD Rock Mass Quality
< 25 Very poor
25 – 50 Poor
50 – 75 Fair
75 – 90 Good
99 – 100 Excellent

RQD has been widely accepted as a measure of fracturing degree of the rock mass. his
parameter has been used in the rock mass classification systems, including the RMR and
the Q systems.
Chapter 6 Rock Mass Properties and Classifications 5

6.2 Rock Mass Rating – RMR System

6.2.1 Concept of RMR System (1973, 1989)

The rock mass rating (RMR) system is a rock mass quality classification developed by
South African Council for Scientific and Industrial Research (CSIR), close associated
with excavation for the mining industry (Bieniawski 1973). Originally, this
geomechanics classification system incorporated eight parameters. The RMR system in
use now incorporates five basic parameters below.

(a) Strength of intact rock material: Uniaxial compressive strength is preferred. For
rock of moderate to high strength, point load index is acceptable.

(b) RQD: RQD is used as described before.

(c) Spacing of joints: Average spacing of all rock discontinuities is used.

(d) Condition of joints: Condition includes joint aperture, persistence, roughness, joint
surface weathering and alteration, and presence of infilling.

(e) Groundwater conditions: It is to account for groundwater inflow in excavation


stability.

Table 6.2.1a is the RMR classification updated in 1989. Part A of the table shows the
RMR classification with the above 5 parameters. Individual rate for each parameter is
obtained from the property of each parameter. The weight of each parameter has already
considered in the rating, for example, maximum rating for joint condition is 30 while for
rock strength is 15. The overall basic RMR rate is the sum of individual rates.

Influence of joint orientation on the stability of excavation is considered in Part B of the


same table. Explanation of the descriptive terms used is given table Part C. With
adjustment made to account for joint orientation, a final RMR rating is obtained, it can be
also expresses in rock mass class, as shown in Table 6.2.1b. The table also gives the
meaning of rock mass classes in terms of stand-up time, equivalent rock mass cohesion
and friction angle.

RMR was applied to correlate with excavated active span and stand-up time, as shown in
Figure 6.2.1a. This correlation allow engineer to estimate the stand-up time for a given
span and a given rock mass.
Chapter 6 Rock Mass Properties and Classifications 6

Table 6.2.1a Rock mass classification RMR system


(a) Five basic rock mass classification parameters and their ratings
1. Strength of intact Point load strength index (MPa) > 10 4 − 10 2−4 1−2
rock material Uniaxial compressive strength (MPa) > 250 100 − 250 50 − 100 25 − 50 5 − 25 1−5 <1
Rating 15 12 7 4 2 1 0
2. RQD (%) 90 − 100 75 − 90 50 − 75 25 − 50 < 25
Rating 20 17 13 8 3
3. Joint spacing (m) >2 0.6 − 2 0.2 − 0.6 0.06 − 0.2 < 0.06
Rating 20 15 10 8 5
4. Condition of joints not continuous, very rough slightly rough surfaces, slightly rough surfaces, continuous, slickensided continuous joints, soft
surfaces, unweathered, no slightly weathered, highly weathered, surfaces, or gouge <5 mm gouge >5 mm thick, or
separation separation <1 mm separation <1 mm thick, or separation 1−5 mm separation >5 mm
Rating 30 25 20 10 0
5. Groundwater inflow per 10 m tunnel length (l /min), or none < 10 10 − 25 25 − 125 > 125
joint water pressure/major in situ stress, or 0 0 − 0.1 0.1 − 0.2 0.2 − 0.5 > 0.5
general conditions at excavation surface completely dry damp wet dripping flowing
Rating 15 10 7 4 0

(b) Rating adjustment for joint orientations


Strike and dip orientation of joints very favourable favourable fair unfavourable very unfavourable
Rating tunnels 0 −2 −5 − 10 − 12
foundations 0 −2 −7 − 15 − 25
slopes 0 −5 − 25 − 50 − 60

(c) Effects of joint orientation in tunnelling

Dip 0° − 20°
Strike perpendicular to tunnel axis
Strike parallel to tunnel axis
Drive with dip Drive against dip
Dip 45° − 90° Dip 20° − 45° Dip 45° − 90° Dip 20° − 45° Dip 45° − 90° Dip 20° − 45° irrespective of strike
very favourable favourable fair unfavourable very unfavourable fair fair
Chapter 6 Rock Mass Properties and Classifications 7

Table 6.2.1b Rock mass classes determined from total ratings and meaning
RMR Ratings 81 − 100 61 − 80 41 − 60 21 − 40 < 20
Rock mass class A B C D E
Description very good good rock fair rock poor rock very poor
rock rock
Average stand-up 10 year for 6 months 1 week for 10 hours 30 minutes
time 15 m span for 8 m 5 m span for 2.5 m for 0.5 m
span span span
Rock mass cohesion > 400 300 − 400 200 − 300 100 − 200 < 100
(KPa)
Rock mass friction > 45° 35° − 45° 25° − 35° 15° − 25° < 15°
angle

Figure 6.2.1a Stand-up time and RMR quality

6.2.2 Examples of using RMR System

(a) A granite rock mass containing 3 joint sets, average RQD is 88%, average joint
spacing is 0.24 m, joint surfaces are generally stepped and rough, tightly closed and
unweathered with occasional stains observed, the excavation surface is wet but not
dripping, average rock material uniaxial compressive strength is 160 MPa, the tunnel is
excavated to 150 m below the ground where no abnormal high in situ stress is expected.

Selection of RMR parameters and calculation of RMR are shown below:

Rock material strength 160 MPa Rating 12


RQD (%) 88% Rating 17
Joint spacing (m) 0.24 m Rating 10
Condition of joints very rough, unweathered, no separation Rating 30
Groundwater wet Rating 7
RMR 76

The calculated basic RMR is 76. It falls in rock class B which indicates the rock mass is
of good quality.

(b) A sandstone rock mass, fractured by 2 joint sets plus random fractures, average RQD
is 70%, average joint spacing is 0.11 m, joint surfaces are slightly rough, highly
weathered with stains and weathered surface but no clay found on surface, joints are
generally in contact with apertures generally less than 1 mm, average rock material
Chapter 6 Rock Mass Properties and Classifications 8

uniaxial compressive strength is 85 MPa, the tunnel is to be excavated at 80 m below


ground level and the groundwater table is 10 m below the ground surface.

Here, groundwater parameter is not directly given, but given in terms of groundwater
pressure of 70 m water head and overburden pressure of 80 m ground. Since there is no
indication of in situ stress ratio, overburden stress is taken as the major in situ stress as an
approximation.

Joint water pressure = groundwater pressure = 70 m × γw


In situ stress = Overburden pressure = 80 m × γ

Joint water pressure / In situ stress = (70 × 1)/(80× 2.7) = 0.32

Selection of RMR parameters and calculation of RMR are shown below:

Rock material strength 85 MPa Rating 7


RQD (%) 70% Rating 13
Joint spacing (m) 0.11 m Rating 8
Condition of joints slightly rough, highly weathered, separation < Rating 20
1mm
Groundwater water pressure/stress = 0.32 Rating 4
RMR 52

The calculated basic RMR is 52. It falls in rock class C which indicates the rock mass is
of fair quality.

(c) A highly fractured siltstone rock mass, found to have 2 joint sets and many random
fractures, average RQD is 41%, joints appears continuous observed in tunnel, joint
surfaces are slickensided and undulating, and are highly weathered, joint are separated by
about 3-5 mm, filled with clay, average rock material uniaxial compressive strength is 65
MPa, inflow per 10 m tunnel length is observed at approximately 50 litre/minute, with
considerable outwash of joint fillings. The tunnel is at 220 m below ground.

In the above information, joint spacing is not provided. However, RQD is given and
from the relationship between RQD and joint frequency, it is possible to calculate average
joint spacing, with the equation below,
–0.1λ
RQD = 100 e (0.1λ +1)

Joint frequency is estimated to be 20, which gives average joint spacing 0.05 m

Selection of RMR parameters and calculation of RMR are shown below:

Rock material strength 65 MPa Rating 7


RQD (%) 41% Rating 8
Chapter 6 Rock Mass Properties and Classifications 9

Joint spacing (m) 0.05 m Rating 5


Condition of joints continuous, slickensided, separation 1-5mm Rating 10
Groundwater inflow = 50 l/min Rating 4
RMR 34

The calculated basic RMR is 34. It falls in rock class D which indicates the rock mass is
of poor quality.

Judgement often is needed to interpret the information given in the geological and
hydrogeological investigation reports and in the borehole logs to match the descriptive
terms in the RMR table. Closest match and approximation is to be used to determine
each of the RMR parameter rating.

6.2.3 Extension of RMR – Slope Mass Rating (SMR)

The slope mass rating (SMR) is an extension of the RMR system applied to rock slope
engineering. SMR value is obtained by adjust RMR value with orientation and
excavation adjustments for slopes, i.e.,

SMR = RMR + (F1⋅F2⋅F3) + F4

where F1 = (1 - sin A)2


and A = angle between the strikes of the slope and the joint = |αj - αs|.

F2 = (tan βj)2
B = joint dip angle = βj.
For topping, F2 = 1.0

Value of F1, F2 and F3 are given in Table 6.2.3a. Table 6.2.3b gives the classification
category of rock mass slope. Details on rock slope analysis and engineering including
excavation methods and support and stabilisation will be covered in a later chapter
dealing slope engineering.
Chapter 6 Rock Mass Properties and Classifications 10

Table 6.2.3a Adjustment rating of F1, F2, F3 and F4 for joints


Very Very
Joint Orientation Favourable Fair Unfavourable unfavourable
favourable
P |αj - αs| >30 30~20 20~10 10~5 <5
T |(αj - αs) - 180| >30 30~20 20~10 10~5 <5
F1 (for P & T) 0.15 0.40 0.70 0.85 1.00
P |βj| <20 20~30 30~35 35~45 >45
F2 (for P) 0.15 0.40 0.70 0.85 1.00
F2 (for T) 1.00 1.00 1.00 1.00 1.00
P βj - βs >10 10~0 0 0~-10 <-10
T βj + βs <110 110~120 >120 -- --
F3 (for P & T) 0 -6 -25 -50 -60
Method Natural slope Presplitting Smooth blasting Blasting/Ripping Deficient blasting
F4 +15 +10 +8 0 -8

Table 6.2.3a Classification of Rock Slope according to SMT


SMR Class Description Stability Failure Support
81~100 I Very good Completely stable None None
61~80 II Good Stable Some blocks Spot
Some joints or
41~60 III Fair Partially stable Systematic
many wedges
Palnar or large Important /
21~40 IV Poor Unstable wedges Corrective
Completely Large wedges or
0~20 V Very Poor Re-excavation
unstable circular failure
Chapter 6 Rock Mass Properties and Classifications 11

6.3 Rock Tunnel Quality Q-System

6.3.1 Concept of the Q-System

The Q-system was developed as a rock tunnelling quality index by the Norwegian
Geotechnical Institute (NGI) (Barton et al 1974). The system was based on evaluation
of a large number of case histories of underground excavation stability, and is an index
for the determination of the tunnelling quality of a rock mass. The numerical value of
this index Q is defined by:

RQD Jr Jw
Q=
Jn Ja SRF

RQD is the Rock Quality Designation measuring the fracturing degree. Jn is the joint set
number accounting for the number of joint sets. Jr is the joint roughness number
accounting for the joint surface roughness. Ja is the joint alteration number indicating
the degree of weathering, alteration and filling. Jw is the joint water reduction factor
accounting for the problem from groundwater pressure, and SRF is the stress reduction
factor indicating the influence of in situ stress.

Q value is considered as a function of only three parameters which are crude measures of:
(a) Block size: RQD / Jn
(b) Inter-block shear strength Jr / Ja
(c) Active stress Jw / SRF

Parameters and rating of the Q system is given in Table 6.3.1a. The classification
system gives a Q value which indicates the rock mass quality, shown in Table 6.3.1b.

Q value is applied to estimate the support measure for a tunnel of a given dimension and
usage, as shown in Figure 6.3.1a. Equivalent dimension is used in the figure and ESR is
given in Table 6.3.1c.

Excavation span, diameter or height (m)


Equivalent dimension, De =
Excavation Support Ratio (ESR)
Chapter 6 Rock Mass Properties and Classifications 12

Table 6.3.1a Rock mass classification Q system


1. Rock Quality Designation RQD
A Very Poor 0 – 25
B Poor 25 – 50
C Fair 50 – 75
D Good 75 – 90

Note: (i) Where RQD is reported or measured as ≤ 10 (including 0), a nominal value of 10 is
E Excellent 90 – 100

used to evaluate Q. (ii) RQD interval of 5, i.e., 100, 95, 90, etc., are sufficiently
accurate.
2. Joint Set Number Jn
A Massive, no or few joints 0.5 – 1
B One joint set 2
C One joint set plus random joints 3
D Two joint set 4
E Two joint set plus random joints 6
F Three joint set 9
G Three joint set plus random joints 12
H Four or more joint sets, heavily jointed 15

Note: (i) For intersections, use (3.0 × Jn). (ii) For portals, use (2.0 × Jn).
J Crushed rock, earthlike 20

3. Joint Roughness Number Jr


(a) Rock-wall contact, and (b) Rock wall contact before 10 cm shear
A Discontinuous joints 4
B Rough or irregular, undulating 3
C Smooth, undulating 2
D Slickensided, undulating 1.5
E Rough or irregular, planar 1.5
F Smooth, planar 1.0
G Slickensided, planar 0.5
Note: (i) Descriptions refer to small and intermediate scale features, in that order.
(c) No rock-wall contact when sheared
H Zone containing clay minerals thick enough to prevent rock-wall contact 1.0

Note: (ii) Add 1.0 if the mean spacing of the relevant joint set ≥ 3 m. (iii) Jr = 0.5 can be used
J Sandy, gravelly or crushed zone thick enough to prevent rock-wall contact 1.0

for planar slickensided joints having lineations, provided the lineations are oriented for
minimum strength.
4. Joint Alteration Number φr approx. Ja
(a) Rock-wall contact (no mineral fillings, only coatings)
A Tight healed, hard, non-softening, impermeable filling, i.e., – 0.75
quartz or epidote
B Unaltered joint walls, surface staining only 25 – 35° 1.0
C Slightly altered joint walls. Non-softening mineral coating, 25 – 30° 2.0
sandy particles, clay-free disintegrated rock, etc.
D Silty- or sandy-clay coatings, small clay fraction (non- 20 – 25° 3.0
softening)
E Softening or low friction mineral coatings, i.e., kaolinite or 8 – 16° 4.0
mica. Also chlorite, talc, gypsum, graphite, etc., and small
Chapter 6 Rock Mass Properties and Classifications 13

quantities of swelling clays

(b) Rock wall contact before 10 cm shear (thin mineral fillings)


F Sandy particles, clay-free disintegrated rock, etc. 25 – 30° 4.0
G Strongly over-consolidated non-softening clay mineral 16 – 24° 6.0
fillings (continuous, but < 5 mm thickness)
H Medium or low over-consolidated softening clay mineral 12 – 16° 8.0
fillings (continuous, but < 5 mm thickness)
J Swelling-clay fillings, i.e., montmorillonite (continuous, but 6 – 12° 8 – 12
< 5 mm thickness). Value of Ja depends on percent of
swelling clay size particles, and access to water, etc.
(c) No rock-wall contact when sheared (thick mineral fillings)
K, L, Zones or bands of disintegrated or crushed rock and clay 6 – 24° 6, 8, or
M (see G, H, J for description of clay condition) 8 – 12
N Zones or bands of silty- or sandy-clay, small clay fraction - 5
(non-softening)
O, P, Thick, continuous zones or bands of clay (see G, H, J for 6 – 24° 10, 13, or
R clay condition description) 13 – 20

5. Joint Water Reduction Factor Water pressure Jw


A Dry excavation or minor inflow, i.e., < 5 l/min < 1 (kg/cm2) 1.0
locally
B Medium inflow or pressure, occasional outwash of 1 – 2.5 0.66
joint fillings
C Large inflow or high pressure in competent rock 2.5 – 10 0.5
with unfilled joints
D Large inflow or high pressure, considerable outwash 2.5 – 10 0.33
of joint fillings
E Exceptionally high inflow or water pressure at > 10 0.2 – 0.1
blasting, decaying with time
F Exceptionally high inflow or water pressure > 10 (kg/cm2) 0.1 – 0.05
continuing without noticeable decay
Note: (i) Factors C to F are crude estimates. Increase Jw if drainage measures are installed.
(ii) Special problems caused by ice formation are not considered.

6. Stress Reduction Factor SRF


(a) Weakness zones intersecting excavation, which may cause loosening of rock mass when
tunnel is excavated
A Multiple occurrences of weakness zones containing clay or chemically 10
disintegrated rock, very loose surrounding rock (any depth)

(depth of excavation ≤ 50 m)
B Single weakness zone containing clay or chemically disintegrated rock 5

C Single weakness zone containing clay or chemically disintegrated rock 2.5

Multiple shear zones in competent rock (clay-free) (depth of excavation ≤


(depth of excavation > 50 m)
D 7.5

Single shear zone in competent rock (clay-free) (depth of excavation ≤ 50


50 m)
E 5
m)
F Single shear zone in competent rock (clay-free) (depth of excavation > 50 2.5
m)
G Loose, open joint, heavily jointed (any depth) 5
Note: (i) Reduce SRF value by 25-50% if the relevant shear zones only influence but not
intersect the excavation.
Chapter 6 Rock Mass Properties and Classifications 14
Chapter 6 Rock Mass Properties and Classifications 15

(b) Competent rock, rock stress problems σc / σ1 σθ / σc SRF


H Low stress, near surface, open joints > 200 < 0.01 2.5
J Medium stress, favourable stress condition 200 – 10 0.01 – 1
0.03
K High stress, very tight structure. Usually 10 – 5 0.3 – 0.4 0.5 – 2
favourable to stability, may be unfavourable to wall
stability
L Moderate slabbing after > 1 hour in massive rock 5–3 0.5 - 0.65 5 – 50
M Slabbing and rock burst after a few minutes in 3–2 0.65 – 1 50 – 200
massive rock
N Heavy rock burst (strain-burst) and immediate <2 >1 200 – 400

Note: (ii) For strongly anisotropic virgin stress field (if measured): when 5 ≤ σ1 / σ3 ≤ 10, reduce
dynamic deformation in massive rock

σc to 0.75 σc; when σ1 / σ3 > 10, reduce σc to 0.5 σc; where σc is unconfined
compressive strength, σ1 and σ3 are major and minor principal stresses, and σθ is
maximum tangential stress (estimated from elastic theory). (iii) Few cases records
available where depth of crown below surface is less than span width. Suggest SRF

σθ / σc
increase from 2.5 to 5 for such cases (see H).
(c) Squeezing rock: plastic flow in incompetent rock under the SRF
influence of high rock pressure
O Mild squeezing rock pressure 1–5 5 – 10
P Heavy squeezing rock pressure 5 10 – 20

strength can be estimated from Q = 7 γ Q1/3 (MPa), where γ = rock density in g/cm3.
Note: (vi) Cases of squeezing rock may occur for depth H > 350 Q1/3. Rock mass compressive

(d) Swelling rock: chemical swelling activity depending on presence of water SRF
R Mile swelling rock pressure 5 – 10
S Heavy swell rock pressure 10 – 15
Note: Jr and Ja classification is applied to the joint set or discontinuity that is least favourable for
stability both from the point of view of orientation and shear resistance.

Table 6.3.1b Rock mass quality rating according to Q values


Q-value Class Rock mass quality
400 ~ 1000 A Exceptionally Good
100 ~ 400 A Extremely Good
40 ~ 100 A Very Good
10 ~ 40 B Good
4 ~ 10 C Fair
1~4 D Poor
0.1 ~ 1 E Very Poor
0.01 ~ 0.1 F Extremely Poor
0.001 ~ 0.01 G Exceptionally Poor
Chapter 6 Rock Mass Properties and Classifications 16

Figure 6.3.1a Support design based on Q value

Table 6.3.1c Excavation Support Ratio (ESR) for various tunnel categories

Excavation Category ESR


A Temporary mine openings. 3–5
Permanent mine openings, water tunnels for hydro-electric
B 1.6
projects, pilot tunnels, drifts and headings for large excavations.
Storage rooms, water treatment plants, minor road and railway
C tunnels, surge chambers and access tunnels in hydro-electric 1.3
project.
Underground power station caverns, major road and railway
D 1.0
tunnels, civil defense chamber, tunnel portals and intersections.
Underground nuclear power stations, railway stations, sports and
E 0.8
public facilities, underground factories.

6.3.2 Examples of Using the Q-System

(a) A granite rock mass containing 3 joint sets, average RQD is 88%, average joint
spacing is 0.24 m, joint surfaces are generally stepped and rough, tightly closed and
unweathered with occasional stains observed, the excavation surface is wet but not
dripping, average rock material uniaxial compressive strength is 160 MPa, the tunnel is
excavated to 150 m below the ground where no abnormal high in situ stress is expected.

Selection of Q parameters and calculation of Q-value are shown below:

RQD 88% RQD 88


Joint set number 3 sets Jn 9
Joint roughness number rough stepped (⇒undulating) Jr 3
Joint alteration number unaltered, some stains Ja 1
Joint water factor wet only (dry excavation or minor inflow) Jw 1
Stress reduction factor σc/σ1 = 160/(150×0.027) = 39.5 SRF 1
Q (88/9) (3/1) (1/1) 44

The calculated Q-value is 29, and the rock mass is classified as good quality.

(b) A sandstone rock mass, fractured by 2 joint sets plus random fractures, average RQD
is 70%, average joint spacing is 0.11 m, joint surfaces are slightly rough, highly
weathered with stains and weathered surface but no clay found on surface, joints are
Chapter 6 Rock Mass Properties and Classifications 17

generally in contact with apertures generally less than 1 mm, average rock material
uniaxial compressive strength is 85 MPa, the tunnel is to be excavated at 80 m below
ground level and the groundwater table is 10 m below the ground surface.

Selection of Q parameters and calculation of Q-value are shown below:

RQD 70% RQD 70


Joint set number 2 sets plus random Jn 6
Joint roughness number slightly rough (⇒rough planar) Jr 1.5
Joint alteration number highly weathered only stain, (altered non- Ja 2
softening mineral coating)
Joint water factor 70 m water head = 7 kg/cm2 = 7 bars Jw 0.5
Stress reduction factor σc/σ1 = 85/(80×0.027) = 39.3 SRF 1
Q (70/6) (1.5/2) (0.5/1) 4.4

The calculated Q-value is 4.4, and the rock mass is classified as fair quality.

(c) A highly fractured siltstone rock mass, found to have 2 joint sets and many random
fractures, average RQD is 41%, joints appears continuous observed in tunnel, joint
surfaces are slickensided and undulating, and are highly weathered, joint are separated by
about 3-5 mm, filled with clay, average rock material uniaxial compressive strength is 65
MPa, inflow per 10 m tunnel length is observed at approximately 50 litre/minute, with
considerable outwash of joint fillings. The tunnel is at 220 m below ground.

Selection of Q parameters and calculation of Q-value are shown below:

RQD 41% RQD 41


Joint set number 2 sets plus random Jn 6
Joint roughness number slickensided and undulating Jr 1.5
Joint alteration number highly weathered filled with 3-5 mm clay Ja 4
Joint water factor large inflow with considerable outwash Jw 0.33
Stress reduction factor σc/σ1 = 65/(220×0.027) = 11 SRF 1
Q (41/6) (1.5/4) (0.33/1) 0.85

The calculated Q-value is 0.85, and the rock mass is classified as very poor quality.

Again, judgement is frequently needed to interpret the descriptions given in the geological
and hydrogeological investigation reports and in the borehole logs to match the
descriptive terms in the Q table. Closest match and approximation is to be used to
determine each of the Q parameter rating.

6.3.3 Extension of Q-System – QTBM for Mechanised Tunnelling


Chapter 6 Rock Mass Properties and Classifications 18

Q-system was extended to a new QTBM system for predicting penetration rate (PR) and
advance rate (AR) for tunnelling using tunnel boring machine (TBM) in 1999 (Barton
1999). The method is based on the Q-system and average cutter force in relations to the
appropriate rock mass strength. Orientation of joint structure is accounted for, together
with the rock material strength. The abrasive or nonabrasive nature of the rock is
incorporated via the cutter life index (CLI). Rock stress level is also considered. The new
parameter QTBM is to estimate TBM performance during tunnelling.

The components of the QTBM are as follows:

RQD0 Jr Jw 209 σm 20 q σθ
QTBM =
Jn Ja SRF F10 CLI 20 5

ratings are the same parameters in the original Q-system, σm is the rock mass strength
where RQD0= RQD (%) measured in the tunnelling direction, Jn, Jr, Ja, Jw, and SRF

(MPa) estimated from a complicated equation including the Q-value measured in the

life index, q is the quartz content (%) in rock mineralogy, and σθ is the induced biaxial
tunnel direction, F is the average cutter load (ton) through the same zone, CLI is the cutter

stress (MPa) on tunnel face in the same zone. The constants 20 in the σm term, 20 in the
CLI term and 5 in the σθ term are normalising constants.

The experiences on the application of QTBM varies between projects. Example of using
the QTBM is given in Figure 6.2.3a. It appears that the correlation between QTBM and
Advanced Rate is not consistent and varies with a large margin.

Rock mass classification systems, including RMR and Q, when developed, were intended
to classify rock mass quality to arrive a suitable support design. The systems were not
meant for the design of excavation methodology. In general, with increasing of rock
mass quality, penetration decreases. However, very poor rock mass does not facilitate
penetration. Parameters in those rock mass classifications were related to support design,
they were not selected to describe rock mass boreability. Although QTBM has added a
number of parameters to reflect cutting force and wear, the emphasis is obviously not be
justified. The original rock mass classifications are independent of TBM characteristics,
while penetration however is a result of interaction between rock mass properties and
TBM machine parameters (Zhao 2006).
Chapter 6 Rock Mass Properties and Classifications 19

6.4 Geological Strength Index GSI System and Others

6.4.1 GSI System

The Geological Strength Index (GSI) was introduced by Hoek in 1994. It was aimed to
estimate the reduction in rock mass strength for different geological conditions. This
system is presented in Tables 6.4.1a. The system gives a GSI value estimated from rock
mass structure and rock discontinuity surface condition. The direct application of GSI
value is to estimate the parameters in the Hoek-Brown strength criterion for rock masses.
Although it was not aimed at to be a rock mass classification, the GSI value does in fact
reflect the rock mass quality.

GSI system has been modified and updated in the recent years, mainly to cover more
complex geological features, such as sheared zones.

The use of GSI requires careful examination and understanding of engineering geological
features of the rock mass. Rock mass structure given in the chart is general description
and there ma ybe many cases that does not directly match the description.

In general, the following equivalent between rock mass structural descriptions of blocky
to the block size description is suggested below. However, simple block size description
does not include geological structural features, such as folds and shear zones.

GSI Description ISRM Designation Jv, joints/m3 RQD, %


Blocky Medium to large blocks < 10 90 ~ 100
Very block Small to medium blocks 10 – 30 60 ~ 90
Blocky/Folded/Faulted Very small to small blocks > 30 30 ~ 60
Crushed Crushed rock > 60 < 30

GSI does not include the parameter of rock strength, as GSI was initiated to be a tool to
estimate rock mass strength with the Hoek-Brown strength criterion. In the Hoek-
Brown criterion, rock material uniaxial strength is used as a base parameter to estimate
rock mass uniaxial strength as well as triaxial strengths of rock material and rock mass.

The use of GSI to estimate rock mass strength is given later in the section dealing with
rock mass strength.

GSI system dis not suggest a direct correlation between rock mass quality and GSI value.
However, it is suggested that GSI can be related to RMR by GSI = RMR – 5, for
reasobale good quality rock mass. An approximate classification of rock mass quality
and GSI is therefore suggested in Table 6.4.1b, base on the correlation between RMR and
GSI.
Chapter 6 Rock Mass Properties and Classifications 20

Table 6.4.1a Geological Strength Index (GSI)

GEOLOGICAL STRENGTH INDEX (GSI)

VERY GOOD – very rough, fresh, un-

thered surfaces with compact coating

VERY POOR – Slickensided, highly


GOOD – rough, slightly weathered,

POOR – Slickensided, highly wea-

weathered, surfaces with soft clay


FAIR – Smooth, moderately wea-
According to rock mass structure and

JOINT SURFACE CONDITION

or fillings or angular fragments.


discontinuity surface conditions observed

thered, and altered surfaces


on the rock mass at site, select the

weathered Joint surfaces


appropriate box in this chart. Estimate the
average value of the GSI from the

stained joint surfaces


contours.

coating or filling
ROCK MASS STRUCTURE ⇒ Decreasing of Surface Quality ⇒
BLOCKY – very well interlocked 80
undisturbed rock mass

⇐ Decreasing Interlocking of Rock Blocks ⇐


consisting of cubical blocks
70
formed by three orthogonal joint
sets
60

VERY BLOCKY – interlocked,


partially disturbed rock mass 50
with multi-faced angular blocks
formed by for or more joint sets.
40

BLOCKY/FOLDED – folded and


faulted with many intersecting
discontinuties forming angular 30
blocks.

CRUSHED – poorly interlocked, 20


heavily broken rock mass with a
mixture of angular and rounded
blocks. 10

Table 6.4.1b Rock mass classes determined from GSI


GSI Value 76 − 95 56 − 75 36 − 55 21 − 35 < 20
Rock Mass Quality Very good Good Fair Poor Very poor
Chapter 6 Rock Mass Properties and Classifications 21

6.4.2 Examples of Using the GSI System

Examples of estimating GSI is given below, with the same rock masses used previously to
estimate RMR and Q.

(a) Granite rock mass containing 3 joint sets, average RQD is 88%, average joint spacing
is 0.24 m, joint surfaces are generally stepped and rough, tightly closed and unweathered
with occasional stains observed, the excavation surface is wet but not dripping, average
rock material uniaxial compressive strength is 160 MPa, the tunnel is excavated to 150 m
below the ground where no abnormal high in situ stress is expected.

Refer to the GSI chart, Rock Mass Structure for the above granite is blocky, and Joint
Surface Condition is very good. Therefore GSI is 75±5. The rock mass is classified as
good to very good quality.

(b) A sandstone rock mass, fractured by 2 joint sets plus random fractures, average RQD
is 70%, average joint spacing is 0.11 m, joint surfaces are slightly rough, highly
weathered with stains and weathered surface but no clay found on surface, joints are
generally in contact with apertures generally less than 1 mm, average rock material
uniaxial compressive strength is 85 MPa, the tunnel is to be excavated at 80 m below
ground level and the groundwater table is 10 m below the ground surface.

Refer to the GSI chart, Rock Mass Structure for the above sandstone is very blocky, and
Joint Surface Condition is fair to poor. Therefore GSI is 40±5. The rock mass is
classified as fair quality.

(c) A highly fractured siltstone rock mass, found to have 2 joint sets and many random
fractures, average RQD is 41%, joints appears continuous observed in tunnel, joint
surfaces are slickensided and undulating, and are highly weathered, joint are separated by
about 3-5 mm, filled with clay, average rock material uniaxial compressive strength is 65
MPa, inflow per 10 m tunnel length is observed at approximately 50 litre/minute, with
considerable outwash of joint fillings. The tunnel is at 220 m below ground.

Refer to the GSI chart, Rock Mass Structure for the above siltstone is
blocky/folded/faulted, and Joint Surface Condition is very poor. Therefore GSI is 20±5.
The rock mass is classified as very poor to poor quality.

It is advised that while selecting an average value of GSI, it is perhaps better to select a
range of the GSI value for that rock mass.

Summary of RMR, Q and GSI from the above three examples are given below,

RMR Quality Q Quality GSI Quality


(a) Granite rock mass 76 G 29 G 75 G
(b) Sandstone rock mass 52 F 4.4 F 40 F
(c) Siltstone rock mass 34 P 0.85 VP 20 VP
Chapter 6 Rock Mass Properties and Classifications 22

6.4.3 Correlation and Comparison between Q, RMR and GSI

Correlation between Q and RMR are found to be,

RMR = 9 lnQ + A

A varies between 26 and 62, and average of A is 44. Figure 6.4.3a shows the
comparison and correlation between RMR and Q.

Figure 6.4.3a Correlation between RMR and Q values.

Several other correlation equations have been proposed, including RMR = 13.5 logQ +
43. They are all in the general form of semi-log equation.

For generally competent rock masses with GSI > 25, the value of GSI can be related to
Rock Mass Rating RMR value as,

GSI = RMR – 5

RMR is the basic RMR value by setting the Groundwater rating at 15 (dry), and without
adjustment for joint orientation. For very poor quality rock masses, the value of RMR
is very difficult to estimate and the correlation between RMR and GSI is no longer
reliable. Consequently, RMR classification should not be used for estimating the GSI
values for poor quality rock masses.

It should be noted that each classification uses a set of parameters that are different from
other classifications. For that reason, estimate the value of one classification from
another is not advisable.

6.4.3 Other Classification Systems

Several other classification approaches have been proposed. In section, a few will be
briefly discussed due to their unique application in certain aspect.

(a) Rock Mass Number, N

Rock Mass Number (N) is the rock mass quality Q value when SRF is set at 1 (i.e.,
normal condition, stress reduction is not considered). N can be computed as,

N = (RQD/Jn) (Jr/Ja) (Jw)


Chapter 6 Rock Mass Properties and Classifications 23

This system is used because the difficult in obtaining SRF in the Q-system. It has been

assign to SRF cover too great range. For example, SRF = 1 for σc/σ1 = 10~200, i.e., for a
noticed that SRF in the Q-system is not sensitive in rock engineering design. the value

rock with σc = 50 MPa, in situ stresses of 0.25 to 5 MPa yield the same SRF value.
The importance of in situ stress on the stability of underground excavation is
insufficiently represented in the Q-system.

Another application of N number is to the rock squeezing condition. Squeezing has


been noted in the Q-system but is not sufficiently dealt, due to the special behaviour and
nature of the squeezing ground. The use of N in squeezing rock mass classification will
be presented in a later section in this chapter.

(b) Rock Mass Index, RMi

Rock Mass Index is proposed as an index characterising rock mass strength as a


construction material. It is calculated by the following equation,

RMi = σc Jp

where σc is the uniaxial compressive strength of the intact rock material, and Jp is the
jointing parameter accounting for 4 joint characteristics, namely, joint density (or block
size), joint roughness, joint alteration and joint size. Jp is in fact a reduction factor
representing the effects of jointing on the strength of rock mass. Jp = 1 for a intact rock,
Jp = 0 for a crushed rock masses.
Chapter 6 Rock Mass Properties and Classifications 24

6.5 Rock Mass Strength and Rock Mass Quality

6.5.1 Strength of Rock Mass

As discussed earlier, strength and deformation properties of a rock mass are much
governed by the existence of joints. In another word, the mechanical properties of a
rock mass is also related to the quality of the rock mass. In general, a rock mass of good
quality (strong rock, few joints and good joint surface quality) will have a higher strength
and high deformation modulus than that of a poor rock mass.

6.5.2 Hoek-Brown Strength Criterion of Rock Mass

Hoek and Brown criterion discussed in Chapter 4 is not only for rock materials. It is
also applicable to rock masses (Figure 6.5.2a). The Hoek-Brown criterion for rock mass
is described by the following equation:

σ1 σ σ
σci σci σci
= 3 + ( mb 3 + s)a

or

σ1 = σ3 + (mb σ3 σci + s σci2)a

Figure 6.5.2a Applicability of Hoek-Brown criterion for rock material and rock masses.

The equation above is the generalised Hoek-Brown criterion of rock mass. The Hoek-
Brown criterion for intact rock material is a special form of the generalised equation when
s =1 and a = 0.5. For intact rock, mb becomes mi, i.e.,

σ1 σ σ
σci σci σci
= 3 + ( mi 3 + 1)0.5

Note in the Hoek-Brown criterion, σci is consistently referred to the uniaxial compressive
strength of intact rock material in the Hoek-Brown criterion for rock material and for rock
mass.

In the generalised Hoek-Brown criterion, σ1 is the strength of the rock mass at a confining
pressure σ3. σci is the uniaxial strength of the intact rock in the rock mass. Parameter a is
generally equal to 0.5. Constants mb and s are parameters that changes with rock type and
rock mass quality. Table 6.5.2a gives an earlier suggestion of mb and s values.
Chapter 6 Rock Mass Properties and Classifications 25

Table 6.5.2a Relation between rock mass quality and Hoek.Brown constants
Hoek-Brown Failure Carbonate Lithified Arenaceous Fine grained Coarse grained
Criterion rocks with well argillaceous rocks with polyminerallic polyminerallic

σ1/σc = σ3/σc + (mb


developed rocks strong crystals igneous igneous and

σ3/σc + s)0.5
crystal (mudstone, and poorly crystalline metamorphic
cleavage siltstone, shale, developed rocks crystalline
(dolomite, slate) (normal crystal (andesite, rocks
limestone, to cleavage) cleavage dolerite, (gabbro,
marble) (sandstone, basalt, gneiss, granit,
quartzite) rhyolite) diorite)
Intact rock material
Laboratory size mi = 7.0 mi = 10.0 mi = 15.0 mi = 17.0 mi = 25.0
specimens free from s = 1.0 s = 1.0 s = 1.0 s = 1.0 s = 1.0
joints
RMR = 100 ,Q = 500
Very good quality
rock mass
Tightly interlocking mb = 3.5 mb = 5.0 mb = 7.5 mb = 8.5 mb = 12.5
undisturbed rock with s = 0.1 s = 0.1 s = 0.1 s = 0.1 s = 0.1
unweathered joints
spaced at 3 m
RMR = 85, Q = 100
Good quality rock
mass
Fresh to slightly
mb = 0.7 mb = 1.0 mb = 1.5 mb = 1.7 mb = 2.5
weathered rock,
s = 0.004 s = 0.004 s = 0.004 s = 0.004 s = 0.004
slightly disturbed with
joints spaced at 1 to 3
m
RMR = 65, Q = 10
Fair quality rock mass
Several sets of
mb = 0.14 mb = 0.20 mb = 0.30 mb = 0.34 mb = 0.50
moderately weathered
s = 0.0001 s = 0.0001 s = 0.0001 s = 0.0001 s = 0.0001
joints spaced at 0.3
to 1 m
RMR = 44, Q = 1.0
Poor quality rock
mass
Numerous weathered
mb = 0.04 mb = 0.05 mb = 0.08 mb = 0.09 mb = 0.13
joints spaced at 30 to
s = 0.00001 s = 0.00001 s = 0.00001 s = 0.00001 s = 0.00001
gouge − clean waste
500 mm with some

rock
RMR = 23, Q = 0.1
Very poor quality
rock mass
Numerous heavily
mb = 0.007 mb = 0.01 mb = 0.015 mb = 0.017 mb = 0.025
weathered joints
s=0 s=0 s=0 s=0 s=0
with gouge − waste
spaced at <30 mm

with fines
RMR = 3, Q = 0.01
Chapter 6 Rock Mass Properties and Classifications 26

Development and application of the Hoek-Brown criterion lead to better definition of the
parameters mb and s. Table 6.5.2b presents the latest definition of mi values for the
intact rock materials, according to different rocks.

Table 6.5.2b Values of constant mi for intact rock in Hoek-Brown criterion


Rock Type Rock Name and mi Values
Granite 32±3 Diorite 25±5 Gabbro 27±3
Intrusive Peridotite (25±5)
Granodiorite 29±3 Dolerite (16±5) Norite 22±5
Igneous

Basalt (16±5)
Extrusive Rhyolite (16±5) Andesite 25±5 Porphyries (20±5)
Diabase (16±5)
Agglomerate
Volcanic Tuff (13±5)
(19±3)
Conglomerate
(4±18) Siltstone 7±2 Mudstone 4±2
Clastic Sandstone 17±4
Sedimentary

Breccia (4±16) Marls (7±2) Shale (6±2)


Crystalline Sparitic limestone Micritic limestone
Carbonate Dolomite (9±3)
limestone (12±3) (10±2) (9±2)
Chemical Gypsum 8±2 Anhydrite 12±2
Organic Coal (8±12) Chalk 7±2
Metamorphic

Foliated Gneiss 28±5 Schist 12±3 Phyllites (7±3) Slate 7±4


Slightly Migmatite (29±3) Amphibolite 26±6
Foliated

(19 ±3)
Non Meta-sandstone
Quartzite 20±3 Hornfels (19±4) Marble 9±3
Foliated

The values in the above table are suggestive. As seen from the table, variation of mi
value for each rock can be as great as 18. If triaxial tests have been conducted, the value
of mi should be calculated from the test results.

Once the Geological Strength Index has been estimated, the parameters which describe
the rock mass strength characteristics, are calculated as follows,

GSI – 100
mb = mi exp ( )
28

For GSI > 25, i.e. rock masses of good to reasonable quality, the original Hoek-Brown
criterion is applicable with,

GSI – 100
s = exp ( )
9

and

a = 0.5
Chapter 6 Rock Mass Properties and Classifications 27

For GSI < 25, i.e. rock masses of very poor quality, s = 0, and a in the Hoek-Brown
criterion is no longer equal to 0.5. Value of a can be estimated from GSI by the
following equation,

GSI
a = 0.65 –
200

Uniaxial compressive strength of the rock mass is the value of σ1 when σ3 is zero. From
the Hoek-Brown criterion, when σ3 = 0, it gives the uniaxial compressive strength as,

σ1 = sa σci

Clearly, for rock masses of very poor quality, the uniaxial compressive strength of the
rock masses equal to zero.

Example of using the Hoek-Brown equation to determine rock mass strength is given
below by the same three examples used for determining the rock mass qualities RMR, Q
and GSI. Calculation in the example uses average values only, although in practice,
range of values should be used to give upper and lower bounds.

(a) Granite rock mass, with material uniaxial strength 150 MPa, mean GSI 75.

From the mi table, mi given for granite is approximately 32.

mb = mi exp[(GSI – 100)/28] = 13.1

s = exp[(GSI – 100)/9] = 0.062

GSI > 25, a = 0.5

The Hoek-Brown equation for the granite rock mass is,

σ1 = σ3 + (mb σ3 σci + s σci2)a

σ1 = σ3 + (1956 σ3 + 1395)0.5

Uniaxial compressive strength of the rock mass is, when σ3 = 0,

σcm = 13950.5 = 37.3 MPa

(b) Sandstone rock mass, with material uniaxial strength 85 MPa, mean GSI 40.

From the mi table, mi given for sandstone is approximately 17.

mb = mi exp[(GSI – 100)/28] = 1.99


Chapter 6 Rock Mass Properties and Classifications 28

s = exp[(GSI – 100)/9] = 0.0013

GSI > 25, a = 0.5

Similarly the uniaxial compressive strength is,

σcm = σ3 + (169 σ3 + 9.4)0.5 = 9.40.5 = 3.1 MPa

(c) Siltstone rock mass, with material uniaxial strength 65 MPa, mean GSI 20.

From the mi table, mi given for siltstone is approximately 7.

mb = mi exp[(GSI – 100)/28] = 0.40

s = exp[(GSI – 100)/9] = 0.00014

GSI < 25, a = 0.65 – (GSI/200) = 0.65 – (20/200) = 0.55

Similarly the uniaxial compressive strength is,

σcm = σ3 + (26 σ3 + 0.59)0.55 = 0.590.55 = 0.75 MPa

6.5.4 Correlations between Rock Mass Quality and Mechanical Properties

Correlations between rock mass strength and rock mass quality are reflected in Table
6.5.2a and the Hoek-Brown criterion relating GSI. The better rock mass quality gives
high rock mass strength. When the rock mass is solid and massive with few joints, the
rock mass strength is close to the strength of intact rock material. When the rock mass
is very poor, i.e., RMR < 23, Q < 0.1, or GSI < 25, the rock mass has very low uniaxial
compressive strength close to zero.

Attempts have also been made to correlated deformation modulus of the rock mass with
rock mass quality.

In situ rock mass modulus (Em) can be estimated from the Q and the RMR systems, in the
equations below,

Em = 25 logQ, for Q > 1

Em = 10 (Q σc/100)1/3

Em = 10(15 logQ+40)/40

Em = 2 RMR − 100, for RMR > 50


Chapter 6 Rock Mass Properties and Classifications 29

Em = 10(RMR−10)/40 for 20 < RMR < 85

The above Em-RMR equations are generally for competent rock mass with RMR greater
than 20. For poor rocks, the equation below has been proposed,

σci 0.5 (GSI/40 – 0.25)


Em = ( ) 10
100

For rock mass with σci < 100 MPa. The equation is obtained by substituting GSI for

is reduced progressively as the value of σci falls below 100. This reduction is based
RMR in the original Em-RMR equation. The Em-GSI equation indicates that modulus Em

upon the reasoning that the deformation of better quality rock masses is controlled by the
discontinuities while, for poorer quality rock masses, the deformation of the intact rock
pieces contributes to the overall deformation process.

6.5.4 Relationship between Hoek-Brown and Mohr-Coulomb Criteria

There is no direct correlation between the linear Mohr-Coulomb Criterion and the non-
linear Hoek-Brown Criterion defined by the two equations. Often, the input for a design

Coulomb parameters c and φ. Attempts have been made by Hoek and Brown to
software or numerical modelling required for rock masses are in terms of Mohr-

estimate c and φ from the Hoek-Brown equation. At the same time, they caution the
user that is a major problem to obtain c and φ from the Hoek-Brown equation.

used directly to obtain parameters c and φ, using for example, plotting the Mohr circle
If a series tests have been conducted on the rock mass, obviously test results should be

and fitting with the best strength envelope, where c and φ can be readily calculated

approach to obtain rock mass Mohr-Coulomb parameters c and φ is by generate a series


Common problems were there is no or limited test results on rock mass. The suggested

σ1–σ3 results by the Hoek-Brown equation. Then plotting the Mohr circle using the
generated σ1–σ3 data and fitting with the best linear envelope, where c and φ can be
readily calculated. Care must be taken when deciding the ‘best’ linear line in fitting the
Mohr circles. It depends on the stress region of the engineering application. For a
tunnel problem, if the depth and stress range is known, the line should be fitting best for
the Mohr circles in that stress region. For a slope problem, the stress region may vary
from 0 to some level of stress, and the fitting a line at low stress level (where the
curvature is the greatest for the non-linear Hoek-Brown strength envelope) is very
sensitive to the stress level. Also, pore pressure needs to be considered as this affects
the effective stress level.
Chapter 6 Rock Mass Properties and Classifications 30

6.6 Squeezing Behaviour of Rock Mass

6.6.1 Squeezing Phenomenon

ISRM (Barla 1995) defines that squeezing of rock is the time dependent large
deformation, which occurs around a tunnel and other underground openings, and is
essentially associated with creep caused by exceeding shear strength. Deformation may
terminate during construction or may continue over a long time period. The degree of
squeezing often is classified to mild, moderate and high, by the conditions below,
(i) Mild squeezing: closure 1-3% of tunnel diameter;
(ii) Moderate squeezing: closure 3-5% of tunnel diameter;
(iii) High squeezing: closure > 5% of tunnel diameter.

Behaviour of rock squeezing is typically represented by rock mass squeezes plastically


into the tunnel and the phenomenon is time dependent. Rate of squeezing depends on
the degree of over-stress. Usually the rate is high at initial stage, say, several
centimetres of tunnel closure per day for the first 1-2 weeks of excavation. Closure rate
reduces with time. Squeezing may continue for years in exceptional cases. Squeezing
may occur at shallow depths in weak and poor rock masses such as mudstone and shale.
Rock masses of competent rock of poor rock mass quality at great depth (under high
cover) may also suffer from squeezing.

6.6.2 Squeezing Estimation by Rock Mass Classification

Based on case studies, squeezing may be identified from rock class classification Q-value
and overburden thickness (H). As shown in Figure 6.6.2a, the division between
squeezing and non-squeezing condition is by a line H = 350 Q1/3, where H is in metres.
Above the line, i.e., H > 350 Q1/3, squeezing condition may occur. Below the line, i.e.,
H < 350 Q1/3, the ground is of generally non-squeezing condition.

Figure 6.6.2a Predicting squeezing ground using Q-value

Another approach predicting squeezing is by using the Rock Mass Number (N). As
discussed in the previous section, N is the Q-value when SRF is set to be 1. The
parameter allow one to separate in situ stress effects from rock mass quality. In situ
stress, which is the external cause of squeezing is dealt separated by considering the
overburden depth. From Figure 6.6.2b, the line separating non-squeezing from
squeezing condition is,

H = (275 N1/3) B–0.1

Where H is the tunnel depth or overburden in metres and B is the tunnel span or
diameter in metres.
Chapter 6 Rock Mass Properties and Classifications 31

Squeezing ground condition is presented by H > (275 N1/3) B–0.1.

It is also possible to characterise the degree of squeezing base on the same figure. Mild
squeezing occurs when (275 N1/3) B–0.1 < H < (450 N1/3) B–0.1 Moderate squeezing
occurs when (450 N1/3) B–0.1 < H < (630 N1/3) B–0.1 High squeezing occurs when H >
(630 N1/3) B–0.1.

Theoretically, squeezing conditions around a tunnel opening can occur when,

σθ > Strength = σcm + Px A/2

where σθ is the tangential stress at the tunnel opening, σcm is the uniaxial compressive
strength of the rock mass, Px is the in situ stress in the tunnel axis direction, and A is a
rock parameter proportion to friction. Squeezing may not occur in hard rocks with high
values of parameter A.

The above equation can be written in the form below for a circular tunnel under
hydrostatic in situ stress field, with overburden stress P, P=γH,

2 P > σcm + P A/2

ISRM classifies squeezing rock mass and ground condition in Table 6.6.2a.

Table 6.6.2a Suggested predictions of squeezing conditions


Degree of Squeezing σθ / σcm (ISRM) σcm / γ H (Barla) σcm / σinsitu (Hoek)
Non squeezing < 1.0 > 1.0 > 0.35
Mild squeezing 1.0 – 2.0 0.4 – 1.0 0.2 – 0.35
Moderate squeezing 2.0 – 4.0 0.2 – 0.4 0.15 – 0.2
High squeezing > 4.0 < 0.2 < 0.15

The prediction equations for squeezing require the measurements of in situ stress and
rock mass strength. Overburden stress can be estimated from the overburden depth and
rock unit weight. Uniaxial compressive strength of the rock mass can be estimated
from the Hoek-Brown criterion with rock mass quality assessment (e.g., GSI).

Studies carried out by Hoek (2000) indicate that squeezing can in fact start at rock mass
strength / in situ stress ratio of 0.3. A prediction curve was proposed by Hoek and
reproduced in Figure 6.6.2c, relating tunnel closure to rock mass strength/in situ stress
ratio. The prediction curve was compared with tunnel squeezing case histories.

Figure 6.6.2c Squeezing prediction curve and comparison with case histories.
Chapter 6 Rock Mass Properties and Classifications 32

6.7 Laboratory and Field Characterisation of Rock Mass

6.7.1 Descriptions of Rock Mass and Matrix

6.7.2 Rock Mass Strength

6.7.3 Rock Mass Deformation Modulus

6.7.4 Groundwater Flow and Permeability

6.7.5 In Situ Stress


6.5.2a

6.1.2a

1
6.1.3a

6.1.3b

2
6.2.1a

6.3.1a

3
6.4.3a

6.5.2a

4
6.6.2a

6.6.2b

5
Case Histories
Strength values considered
reliable
Strength values estimated

6.6.2c

You might also like