Вы находитесь на странице: 1из 8

Chapter 2.

In Situ Rock Stress Determination:


Techniques and Applications

JAMESR. AGGSON
VERNEE. HOOKBR
GENERAL CONSIDERATIONS

ABSOLUTE OR FIELD STRESSES

The design of a mine or any underground opening


in rock is similar to the design of any other structure,
such as a building or a bridge, in one important way.
This similarity lies in the fact that the stability of the
structure is a function of the loads that will be applied
during the structure's expected lifetime. In the case of
an underground opening, the applied load is the product
of the existing in situ stress field and the concentration
of this field that is produced by the extraction of a loadbearing material. Thus, a knowledge of the in situ stress
field is considered a basic requirement in the design
process and essential to the eventual stability of the
underground structure (Duvall, 1976; Obert, 1972;
Merrill and Peterson, 1961; Dahl and von Schonfeldt,
1976; Hedley and Wilson, 1975).
The in situ stress field can be used along with the
physical properties of the rock masses involved, the geometry of the various rock types, the geometry of the
openings involved, and a failure criterion to evaluate
structural stability of a planned mine by theoretical or
numerical techniques. This same information can be
used to investigate ground control problems in existing
mines.
In situ rock stress determinations can be separated
into three categories: the determination of applied
stress, also called the absolute or field stress; the determination of concentrations of the field stress near openings; and the determination of changes in rock stress as
load-bearing material is removed. The applied stress
field is more difficult to determine than are stress
changes. Various instruments and techniques have been
developed to determine applied rock stress and stress
changes in rock.

Theoretical Considerations
In the absence of actual measurements, it is often
assumed that at any point, the vertical stress is equal to
the weight of the overlying material.
where S, is vertical stress, 6 is average weight density of
overburden rock, and h is depth from the surface. This
is normally consistent with actual experience.
The horizontal stresses are assumed to be equal and
related to the vertical stress through the Poisson effect.

where S, is horizontal stress, S, is vertical stress, and v


is Poisson's ratio.
The derivation of Eq. 2 is dependent on lateral confinement of the rock mass. An increasing amount of
information is rapidly becoming available which indicates that the in situ stress field at a particular location
cannot simply be assumed to be a gravity-induced stress
distribution in which the secondary principal stresses in
the horizontal plane are due to confinement and Poisson's effect. Rather they can be a complex function of
gravity loading, regional stresses, structural features, excavation geometries, and variations in surface and/or
possibly basement rock topography.
Furthermore, the underground principal stress distribution is not necessarily vertical and horizontal as is
often assumed. The deviation of the three-dimensional
stress distribution from vertical and horizontal can be
due to a combination of both topographic (Hooker,
Bickel, and Aggson, 1972) and structural features.

Fig. 1. Locations of in situ rock


stress determination sites in
which the authors have been directly or indirectly involved.
Stress ellipses represent measured excess horizontal compressive stress. Site numbers correspond to Tables 1 and 2.
Spotted portions represent s u r face sites and striped portions
represent underground sites. Metric equivalents: mile X 1.609 344
= krn; psi X 6894.757 = Pa.

DESIGN CONTROL
Actual Surface and Underground Stress Values
Fig. 1 shows the locations of in situ rock stress determination sites in which the authors have been directly or indirectly involved. The magnitudes and directions of the excess principal stress (P and Q) in the
horizontal plane at the various locations are indicated
in Fig. 1 and presented in Tables 1 and 2. The term
"excess horizontal stress" is used to describe that compressive stress which is found to exceed the horizontal
stresses that are generated by gravity loading and
Poisson's effect.
All stress distributions shown in Fig. 1 have had the

gravity-induced horizontal stress, as calculated by Eq. 1,


subtracted out. Also, the stress determination sites in
Fig. 1 were carefully selected such that the regional applied stress would be determined rather than the stress
concentrations due to mining o r excavation geometry.
The surface sites shown in Fig. 1 were located in
dimension stone quarries or rock outcrops (Hooker and
Duvall, 1966; Hooker and Johnson, 1969; Hooker and
Duvall, 197 1; Hooker, Aggson, and Bickel, 1974). With
the exception of the site in Nevada, all underground sites
are below any nearby topographic relief (or drainage
level) and thus within a contained rock mass. The site

Table 1. Excess Horlzontal Compressive Stress, Surface Sites

No.

Location

Direction
of P

Magnitude
of P,
psi'

1
2
3
4
5
6
7
8
9
10
11
12
13
14

Lithonia, GA
Douglasville, GA
Mt. Airy. NC
Rapidan. VA
St. Peters, PA
West Chelmsford, MA
Proctor. VT
Barre, VT
Graniteville, MO
St. Cloud, MN
Carthage, MO
Troy, OK
Marble Falls, TX
Green River, WY

N499
N64W
N87F
N 6%
N145
N56%
N 4W
N14"E
N77%
N50E
N29
N84W
N33W
N42%

1639
512
2464
1678
820
2133
1328
1734
3190
2205
1066
1075
2219
415

Site

Magnitude
of Q ,
psi
941
285
1191
1385
335
1113
516
791
1397
1519
777
519
1491
171

Average
depth of
measurement,
ft'

Reference

18.1
1.8
33
8.6
4.8
61.9
1.2
151.2
4.7
4.9
4
4.5
4.7
10

Hooker and Johnson, 1969


Hooker and Johnson, 1969
Hooker and Johnson. 1969
Hooker and Johnson, 1969
Hooker and Johnson, 1969
Hooker and Johnson, 1969
Hooker and Johnson, 1969
Hooker and Johnson, 1969
Hooker and Johnson, 1969
Hooker, Aggson, and Bickel, 1974
Hooker and Johnson, 1969
Hooker and Johnson, 1969
Hooker and Johnson. 1969
Previously unpublished

'Metric equivalents: ft x 0.3048 = m; psi x 6894.757 = Pa.

Table 2. Excess Horlzontal Compressive Stress, Underground Sites


-

Site
No.

Location

15

lmmel mine,
Knoxville, TN
Limestone mine,
Barberton, OH
Mather mine,
Ishpeming, MI
Fletcher mine,
Bunker, MO
Homestake mine,
Lead, SD
Crescent mine,
Wallace. ID
Henderson mine,
~mpire,CO
Sunnyside mine,
Sunnyside, UT
Allied Chemical mine,
Green River, WY
Big Island mine,
Green River, WY
Rainier Mesa,
Nevada test site
Lakeshore mine,
Casa Grande, AZ
Beckley No. 1 mine,
Bolt, WV

16
17
18
19
20
21
22
23
24
25
26
27

Direction
of P

Magnitude
of P,
psi'

Magnitude
of Q,
psi

Depth of
Overburden,
ft '

N589

3007

551

925

N77"E

4000

2500

2300

Obert. 1962

N82W

3822

2937

3200

Aggson, 1972

N17W

3682

1595

lo00

Horino and Aggson, 1976

N38F

2778

1053

6200

Previously unpublished

N27W

6258

4966

5300

Skinner, Waddell, and Connway, 1974

N15W

3398

2283

3127

Hooker, Bickel, and Aggson, 1972

N31W

3718

2898

1060

Previously unpublished

N23W

1781

404

1600

Previously unpublished

N38W

1054

705

850

Previously unpublished

N46"E

972

345

1250

Hooker, Aggson, and Bickel, 1971

N17W

502

160

1570

Bickel and Dolinar, 1976

N695

2973

1466

700

'Metric equivalents: ft x 0.3048 = m; psi x 6894.757 = Pa.

Reference
Previously unpublished

Previously unpublished

1500

UNDERGROUND MINING METHODS HANDBOOK

in Nevada is within Rainier Mesa at the Nevada test site


(Hooker, Aggson, and Bickel, 1971). The location of
the stress determination site was on the eastern edge of
the Rainier Mesa and 350.5 m above its base. The
orientation of the maximum horizontal compressive
stress is N45"E. This direction is roughly parallel to the
long axis of Rainier Mesa, and thus would be the direction in which any regional stress distribution would best
be contained. The minimum secondary horizontal stress
at this site is perpendicular to the long axis of the mesa.
This is the direction in which there is the least containment or transmission of the regional stress field. Subsequent investigations conducted by the Special Projects
Branch of the US Geological Survey have shown that
the maximum horizontal compressive stress within the
mesa generally follows the topographic contour of the
mesa (Ellis and Ege, 1976). The magnitudes and directions at this site are an example of how the field stress
can be modified by topography.
The apparent conflict between data points 9 and 18
is explained by the fact that the data for site 18 was
obtained in an underground borehole that was drilled
vertically into the extremely irregular surface of the
Precambrian igneous basement rock that underlies the
mine under investigation. This basement rock outcrops
in the mine. The determined orientation of the horizontal stresses in this underground outcrop may have been
a function of the site that was selected and its relationship to the surface contours of the igneous basement
rock. The orientation of stresses at site 18 was also
controlled by local structure.
All other horizontal compressive stress distributions
shown in Fig. 1 are believed to represent regional horizontal stresses. Fig. 1 is included in this discussion to
indicate the stress distributions that have been determined. Caution should be exercised when considering
the temptation to extrapolate from these data points to
other areas because geologic structure, discontinuities,
and relative stiffness of various geologic members can
significantly affect the continuity of horizontal secondary
principal stresses in the earth's crust.
INSTRUMENTATION
Introduction
In recent years, a large number of stress determination devices and techniques have been developed or proposed. As complete coverage of all devices and proposed techniques is beyond the scope of this discussion,
the devices and techniques included are those that are
most commonly used and accepted by those in the scientific community who are actively investigating in situ
rock stress throughout the world.
When considering which of the stress determination
techniques is best suited to a particular site or problem
under investigation, attention should be given to the
following three selection factors: soundness of the theoretical basis of each of the available methods; the limitations, practicability, or ease of use of each of the available methods in the particular situation; and the direct
cost of equipment and the time required at current labor
rates.
These three factors cannot be considered independently. For example, the rate at which data can be
gathered using a particular method must be considered
when evaluating both the second and third selection fac-

tors. Data are usually obtained at a slower rate when


using techniques that require the use of epoxy or grout.
Trade-offs between the various selection factors must
also be considered. However, if stress information is
a necessity, reliability of results should be a prime
consideration.
For the purposes of this discussion, the devices and
techniques that are used to determine absolute or applied rock stress are divided into the following groups:
deformation techniques, strain cells, hydraulic methods,
and other methods.
The borehole techniques which will be discussed in
this section on instrumentation are more versatile than
those techniques that are limited to surfaces. Borehole
techniques can be used to investigate pillar loads by
obtaining stress profiles from the outside to the center
of the pillar. Borehole techniques can also be readily
used to investigate stress distribution and concentrations
near underground openings.
Absolute Stresses-Deformation Techniques
Borehole Deformation Gages: Borehole deformation
gages are used in conjunction with the overcoring
method of producing stress relief. The method basically
consists of drilling a pilot borehole, positioning the deformation gage in the pilot borehole, and diamond drilling a concentric borehole over the gage and recording
the change in length of the pilot borehole diameter. The
deformation gage and details of the overcoring process
have all been well-documented (Hooker and Bickel,
1974). The US Bureau of Mines (USBM)-type borehole deformation gage measures three diametral deformations in the same plane during overcoring (Hooker,
Aggson, and Bickel, 1974). The thick-walled cylinders
that are obtained after overcoring can be tested in a
biaxial pressure cell (Becker, 1968) to determine the
anisotropic characteristics of the rock. The core can also
be tested in a triaxial pressure cell at stress levels comparable to in situ levels (Obert, 1964). The unloading
secant elastic properties obtained from the triaxial test
are used in the subsequent calculations because these
properties compensate for nonlinearities in the stress1
strain curve of the rock. The deformation measurements
and the elastic properties can then be combined to calculate the stress distribution in the plane normal to the
borehole (Merrill and Peterson, 1961; Hooker and Johnson, 1969; Obert, Merrill, and Morgan, 1962). If borehole deformation measurements are obtained in three
nonparallel boreholes, the complete three-dimensional
state of stress can be calculated (Panek, 1966). The
USBM-type borehole deformation gage has been evaluated under both field (de la Cruz and Raleigh, 1972;
Ageton, 1967; Merrill, et a]., 1964; Ven-Heerden and
Grant, 1967) and laboratory conditions (Merrill and
Peterson, 196 1; Austin, 1970). The laboratory tests
consisted of overcoring the deformation gage in large
cylinders under known load. Calculated stresses were
within 15% of the applied load. The concrete used in
these experiments contained 38.1-mm ( 1.5-in.) coarse
aggregate. Thus errors of 15% are not unreasonable
because the elastic properties would be highly variable.
Errors of less than 5% were obtained in granite.
The deformation gage has been used successfully in
vertical boreholes up to 24.7 m (81 ft) in depth (Palmer
and Lo, 1976) and in horizontal boreholes up to 41.1 m

DESIGN CONTROL
(135 ft) in depth (Merrill, 1970). The use of this gage
and overcoring method is limited to rock in which a
minimum of 203.2 mm (8 in.) of continuous core can
be recovered (Hooker, Aggson, and Bickel, 1974).
Displacement Rosette: Deformations of a rock mass
which result from the drilling of a borehole can be used
to determine surface stresses and stress changes at a rock
surface. This method, which is called the displacement
rosette method or bndercoring, is accomplished by installing six stainless steel pins across three diameters of
a proposed drill hole in a rock surface and measuring
the diametral distance between the three pairs of pins
with a Whittemore gage which has a dial indicator. A
hole is then drilled concentric to the six pins and the
distance between pairs of pins is remeasured. The
change in length of the diametral distances is used along
with the elastic modulus of the rock to calculate the
surface stress distribution (Hooker, Aggson, and Bickel,
1974). This method is inexpensive; the hole can be
drilled with a bolting machine adapted with a masonry
bit; and stress changes in the rock will continuously be
registered by displacement of the pins. But it is limited
to rock surfaces.
Absolute Stresses-Strain
Cells
Surface Strain Gages: The first attempt to determine
in situ rock stress is believed to have been conducted in
1932 (Lieurance, 1933). In this investigation, 508-mm
(20-in.) strain gages on a rock surface were stressrelieved by a circular ring of jackhammer holes. The
low cost of electrical resistance strain gages makes the
use of such devices attractive when considering in situ
rock stress determinations. The direct strain gage technique consists of bonding electric resistance strain gages
directly to a rock surface or borehole bottom and overcoring. In principle, this technique is comparable to the
doorstopper method discussed in the next section. However, in practice, direct strain gage measurements have
numerous practical disadvantages. Electric resistance
strain gages are extremely sensitive to environmental
conditions such as temperature changes and moisture.
These factors can be minimized by the use of temperature compensation gages and moisture proofing. The
effects of differences in the relative size of the strain
gage and the grain size of the material under investigation have not been completely determined. The direct
strain gage technique appears straightforward and is attractive from a materials cost point of view. However,
the practical problems of temperature compensation,
strain gage size, moisture proofing, the temperature of
drill water, and the quality of the bond between the
gage and the rock make the direct strain gage method of
in situ stress determination the most undesirable of those
methods included in this discussion.
The Doorstopper: The South African Council for
Scientific and Industrial Research (CSIR) developed a
device which has come to be known as the "doorstopper"
(Leeman, 1964). It consists of a strain gage rosette
potted in the base of a low-modulus cylindrical solid.
The doorstopper is glued to the end of the borehole
which has been ground flat and smooth with a flat-faced,
diamond-impregnated bit. The strain gages in the doorstopper are mounted close to the end face and readings
taken before and after overcoring give a measure of the
stress relief. The in situ state of stress within the rock

1501

mass is related to the stress relief measured at the face


of the borehole by the use of stress concentration factors. The elastic properties of the rock are used with
the overcoring data to calculate the stress. Readings in
three boreholes are required to calculate the complete
three-dimensional stress field. Questions have been
raised regarding the stress concentration factors and
the soundness of the theoretical basis (de la Cruz and
Raleigh, 1972). The doorstopper itself is relatively inexpensive and has been used quite extensively by
numerous investigators.
CSIR Triaxial Strain Cell: The CSIR triaxial strain
cell consists of three groups of three strain gages
mounted on the circumference of a tubular body (Leeman, 1969). This device is glued to the wall of a pilot
borehole and overcored in a manner similar to that used
with the borehole deformation gage. The strain changes
induced by overcoring and the elastic properties of the
rock are used to calculate the complete in-situ state of
stress from overcoring. Difficulties can exist with adhesives in wet holes.
CSIRO Hollow Inclusion Cell: The CSIRO hollow inclusion cell is similar in construction and concept to the
CSIR triaxial strain cell (Worotnicki and Walton, 1976).
The nine strain measurements allow for some redundancy of measurement and calculation of the complete
state of stress from measurements in a single borehole.
Photoelastic Gages: The photoelastic strain gage
technique is similar in concept to the doorstopper technique and has similar costs and disadvantages (de la Cruz
and Raleigh, 1972). The photoelastic gage consists of
a birefringent plastic disk which has a reflective backing
(Hawkes, 1968). The photoelastic disk is strained during stress relief overcoring. The strains in the disk can be
interpreted by viewing with polarized light and are then
used to determine the rock stress (Hawkes and Moxon,
1965).
Absolute Stresses-Hydraulic
Methods
Flat Jacks: The flat-jack method of rock stress determination was first described in 1952 (Habid and
Marchand, 1952). This method basically consists of
first mounting two vibrating wire strain gages oriented
to measure in the line of intended stress determination,
and then making an initial strain reading. A slot is cut
between the strain gages, and a thin hydraulic cell called
a flat jack is grouted into the slot. After the grout has
cured, the flat jack is pressurized to a value such that the
strain gages indicate their initial value. The flat-jack
pressure is then considered to be equal to the rock stress
that existed normal to the plane of the flat jack before
the slot was cut. This method has been modified to incorporate strain gages and small hydraulic cells (Panek
and Stock, 1964) in boreholes near the flat jack. The
experimental procedure has also been simplified by the
use of a cutting machine to cut the slot (Rocha, Lopes,
and Da Silva, 1966). The flat-jack technique has the disadvantage of being limited to measurements near the
surface of an underground opening and thus influenced
by the geometry of the opening and resulting stress concentrations. However, it has the advantage of not requiring a knowledge of the elastic properties of the rock.
The averaging effect due to the comparatively large area
of the flat jack can also be a significant advantage. This
method can be used in highly stressed areas or rock in

1502

UNDERGROUND MINING METHODS HANDBOOK

which poor core recovery would eliminate those methods that require drill core recovery.
Hydraulic Fracturing: The hydraulic fracturing
method of rock stress determination was first demonstrated in 1957 (Hubbert and Willis, 1957). Since then,
numerous investigators have examined hydraulic fracturing as a means of in situ rock stress determination
(Scheidegger, 1960; Dunlap, 1963; Kehle, 1964; Raleigh, Healy, and Bredehoeft, 1972; Haimson, et al.,
1974). This method of rock stress determination consists of sealing off a section of borehole with packers and
then increasing the fluid pressure in the section of the
hole until the surrounding rock is fractured. The magnitudes of the stresses acting normal to the borehole are
calculated from the relationships between stress concentrations in the vicinity of the sealed-off section and the
fluid pressures involved during and after rock fracture.
An impression packer, which is pressurized in the hydraulically fractured borehole material, is used to obtain
the orientation of the induced fracture and thus determine the stress orientation. The elastic properties of the
rock are not required to calculate stress magnitudes.
However, a knowledge of the tensile strength of the rock
in the fractured interval is required. The theoretical development of the hydraulic fracturing method of stress
determination is based on the assumption that the drill
hole is in the direction of one of the principal stresses.
In an oilfield or in other flat-lying sedimentary formations this may be a reasonable assumption, but in metamorphic and igneous formations, and especially in the
proximity of mineral deposits where the geology is complex, the assumption may be invalid (Obert and Duvall,
1967).
Due to the fact that the hydraulic fracturing method
of stress determination does not require some type of
stress or strain relief by overcoring, this method is the
only method that has been successful in deep [up to
1920-m (6300-ft)] boreholes (Haimson, 1974).
Other Methods
Core Disking: The phenomenon known as core disking may be used to infer in situ rock stress magnitude.
Core disking is the formation of disks or wafers of relatively uniform thickness which fracture or rupture on
surfaces approximately normal to the axis of the drill
core. Usually, the surfaces of the disks are concaveconvex with the concave side toward the collar of the
hole. Core disking is independent of core diameter and
has been observed in drill cores 1.5 m ( 5 ft) diam down
to less than 25 mm (1 in.) diam. The presence of a
center hole in the core, such as would be required for
overcoring, does not influence the disking process.
Core disking, often called "poker chipping" by diamond drillers, occurs when the in situ stress field is high
in relation to the strength properties of the rock involved. The relationships between the in situ stress field,
the strength properties of the rock, and the stress concentrations caused by the borehole and the kerf of the
drilling bit are rather complex. However, laboratory investigations (Obert and Stephenson, 1965), modeling
studies (Durelli, Obert, and Parks, 1968), and field observation (Hooker, Bickel, and Aggson, 1972) have provided various relationships which allow estimation of
the stress field magnitudes.
The drill core from a vertical hole near the earth's

surface or from an underground borehole near the opening from which it is drilled will disk if the average compressive stress in the plane normal to the borehole is
approximately equal to one-half of the unconfined compressive strength of the rock. At this threshold stress
level, the thickness of the disks that are formed will be
roughly one-fourth of the diameter of the core. As the
magnitude of the stress field increases above the threshold stress, the disks will become thidner (shorter in the
axial direction). At yet higher stress levels, the disks
become very thin and may have sharp edges. Empirical
relationships have been developed which allow stress
field estimation from rock strength parameters and disking which occurs in deep holes under triaxial loading
conditions (Obert and Stephenson, 1965). Core disking, when observed in exploration holes from the surface or in drill holes underground, should be taken as
an indication that stress levels are high relative to rock
strength and that ground control problems may be
anticipated.
Explosive Fracturing: The orientation of the maximum and minimum normal compressive stress in rock
can be determined near a rock surface without the need
of absolute stress measurement by one of the methods
described in the previous sections. This can be done by
drilling a hole to a depth which is below the critical
crater depth and then detonating an explosive charge in
the borehole. If the stress field components are of a
biaxial nature, a fracture will propagate in the direction
of the maximum normal compressive stress. The direction of fracture propagation is determined by the principle of least work. This process has been shown to
consistently produce fractures that are in line with the
measured maximum compressive stress (Hooker, Nicholls, and Duvall, 1964).
Stress Change
Introduction: Investigations into the stability of
underground structures often require the determination
of stress change. Stress change in a rock mass or pillar
can be determined much more readily than absolute
stress levels. The borehole deformation gage, for example, can be placed in a borehole and as stress conditions
change, the change in shape of the borehole, as measured by the gage, can be used to directly calculate stress
change in the plane normal to the hole. The displacement rosette and flat-jack methods of absolute stress determination, which were discussed in previous sections,
can be used quite conveniently to monitor stress change
after absolute determinations have been completed. Specialized instrumentation and techniques have been developed which are intended for use in those situations
where stress change is of concern.
Photoelastic Stress Meter: The photoelastic stress
meter consists of a glass cylinder, a sheet of polaroid
material, a quarter wave plate, and a light source
(Roberts, et al., 1964). This device is cemented into a
borehole in such a manner that a change in stress in the
rock is transmitted to the glass cylinder, which in turn
becomes stressed. Since glass is a material that has
photoelastic properties, fringe patterns develop in the
glass and can be viewed from the collar of the borehole.
The fringe pattern in the glass is thus an indication of
the change in stress normal to the borehole. A device
similar in principle, but without the light source, can

DESIGN CONTROL

1503

also be used (Hiramatsu, Niwa, and Oka, 1957). The


light required to create the fringe patterns is supplied at
the collar of the hole, passed through the glass, and then
reflected back to the collar.
Vibrating Wire Stress Meter: The vibrating wire
stress meter (Anon., 1973) consists of a rigid cylindrical
steel cell with a highly tensioned wire across one diameter. This cylinder is wedged into a borehole, and a zero
reading is taken by electronically plucking the tensioned
wire. The frequency of vibration of the wire, which is
read out on an indicator, is proportional to the tension
in the wire (Hawkes and Hooker, 1974). A change in
rock stress is transmitted to the cylinder and changes the
wire tension. This change is detected by the indicator as
a change in the vibrational frequency of the wire. Indicator reading changes are then directly related to rock
stress changes.
Borehole Pressure Cells: Cylindrical borehole pressure cells and flat borehole pressure cells can be used to
detect changes in rock pressure (Smith, 1972). These
devices are set in boreholes under pressure. Subsequent
changes in fluid pressure can be used to infer rock stress
changes if the response of the cell in the particular rock
medium is known. Flat borehole cells can give directional information, while the cylindrical pressure cells
give directionally averaged information.
Geophysical Techniques: The longitudinal propagation velocity of a spherical wave in rock is a function of
the elastic properties of the rock (Obert and Duvall,
1967). In a linear-elastic material the elastic properties
are independent of the state of stress and propagation
velocity in such a material is a constant. However, no
material is perfectly linear elastic. Thus, the elastic
properties are a function of the applied stress. Since the
elastic properties are a function of the applied stress, so
then is the propagation velocity. This is the basis for the
seismic method of rock stress determination. Attempts to
utilize this method to determine absolute stress levels by
establishing a velocity-stress relationship for various
rock types have not been successful. However, repetitive velocity surveys can be used quite effectively to
monitor stress changes associated with underground excavations (Blake, Leighton, and Duvall, 1974).
Mapping of Structural Geology: Attempts have been
made to estimate both the magnitude and the direction
of in situ stress fields from an analysis of various components of the structural geology. The study of fractures, fracture symmetry, and large-scale deformation
features can indeed lead to an understanding of the
forces that created such features. These forces, however, existed millions of years ago and may have little
relationship to the active applied forces in existence today. It is true that the structural geology of a rock mass
does indeed partially determine stress magnitudes and
directions. It is impossible, however, to predict the
in situ stress distribution from geologic information
alone without a knowledge of the far-field applied
boundary stresses. These far-field forces may be the
same forces associated with the plate tectonics model of
the earth that has gained rapid acceptance by today's
earth scientists.

depends on the stress magnitudes involved, the material


that is being mined, the mining system that is selected,
and various other factors. A theoretical stability analysis
or numerical modeling investigation that is conducted
prior to mining will be severely limited if actual in situ
stress levels are not used as input to the analysis or
model.
The geometry of underground entries and openings
is usually selected on the basis of extraction convenience
and equipment requirements. The opening geometries
may have a large effect on opening stability and ground
control. For example, since most rock is usually much
weaker in tension than in compression, opening geometries which produce tensile stress concentrations under
certain loading conditions (Obert and Duvall, 1967)
should be avoided if in situ stresses are similar to these
loading conditions. Similarly, since stress concentrations
increase as the radius of curvature of a corner decreases
(Obert and Duvall, 1967), openings with sharp corners
are most undesirable in highly stressed ground.
The orientation of openings relative to the stress
field is also a design consideration. Since most in situ
stress fields are of a biaxial nature (one component
larger than the other), entries that are parallel to the
maximum normal compressive stress will tend to require
less support than entries in other directions. This is
because stress concentrations around underground openings do not depend on the stress magnitude parallel to
the opening. Thus, entries that parallel the maximum
normal compressive stress direction will be perpendicular to the least stress and will, therefore, have stress concentrations of lesser magnitudes. Entries that are aligned
with the maximum compressive stress would be best
selected for haulage, ventilation, and other functions
requiring long-term usage and stability.
In sedimentary deposits beam theory is often used to
predict stresses, deflection, and sag, as well as to determine bolting patterns and lengths. To be accurate, an
analysis of this type should include the effects of horizontal in situ stress applied parallel to the axis of the
beam.
All effects of excess in situ stress fields are not negative in an underground mining situation. Pillar strengths,
for example, can actually be increased by the existence
of excess horizontal stress. This is due to the fact that
horizontal stress in the roof and floor can be transmitted
into the pillar and can increase the lateral confinement
applied to the pillar material. This increase in lateral
pressure on the pillar material will significantly increase
axial load-carrying capacity.
If a roof member in an underground mine has the
proper geometry to behave as predicted by beam theory,
sag and tensile failure of the roof member might be
avoided by the presence of horizontal stress of sufficient
magnitude. The stress parallel to the beam axis may
tend to hold the beam entirely in compression.
Complete knowledge of the in situ stress field associated with an underground mine is, of course, not
sufficient to avoid or solve all stability or ground control
problems. It is, however, a necessary portion of the
total input to the overall engineering process.

PRACTICAL APPLICATION
The manner in which the stress field information, once
obtained, is used in the underground mine design process

SUMMARY
The state of stress associated with a planned or existing underground excavation cannot simply be as-

LINDERGROLIND MINING METHODS HANDBOOK


sumed to be totally produced by the gravity load of the
overburden. If the elevation of a n underground excavation is less than the surrounding ground surface elevation, it is highly probable that in situ stresses a r e of
sufficient magnitude t o affect the eventual structural
stability. T h e lack of structural stability will cause a
combination of safety and economic problems. A knowledge of the magnitudes and orientations of the in situ
stress field can b e obtained by using selected instrumentation and techniques discussed in this chapter.

REFERENCES AND BIBLIOGRAPHY


Ageton, R. W., 1967, "Deep Mine Stress Determinations
Using Flat Jack and Borehole Deformation Methods,"
Report of Investigations 6887, US Bureau of Mines.
Aggson, J. R., 1972, "Report on In Situ Determination of
Stresses, Mather Mine, Ishpeming, Michigan," Progress
Report DMRC 10006, US Bureau of Mines.
Anon., 1973, Final Report, Contract No., H0220050, US
Bureau of Mines.
Austin, W. G., 1970, "Development of a Stress Relief
Method with a Three-Dimensional Borehole Deformation
Gage," Reclamation Report REC-OCE-7Q.10, US Bureau
of Mines.
Becker, R. M., 1968, "An Anisotropic Elastic Solution for
Testing Stress Relief Cores," Report of Investigations 7143.
US Bureau of Mines.
Bickel, D. L., and Dolinar, D. R., 1976, "In Situ Determination of Stresses, Lakeshore Mine, Casa Grande, Arizona,"
Progress Report DMRC 10018, US Bureau of Mines.
Blake, W., Leighton, F., and Duvall, W. I., 1974, "Microseismic Techniques for Monitoring the Behavior of Rock
Structures," Bulletin 665, US Bureau of Mines.
Dahl, D. H., and von Schonfeldt, H. A., 1976, "Rock
Mechanics Elements of Coal Mine Design," Monograph
on Rock Mechanics Applications in Mining, W. S. Brown,
S. J. Green, W. A. Hustrulid, eds., AIME, New York,
pp. 31-39.
De la Cruz, R. V., and Raleigh, C. B., 1972, "Absolute
Stress Measurements at the Rangely Anticline, Northwestern Colorado," International Journal of Rock Mechanics and Mining Science, Vol. 9, pp. 625-634.
Dunlap, I. R., 1963, "Factors Controlling the Orientation
and Direction of Hydraulic Fractures," Journal, Institute
of Petroleum Engineers, Vol. 49, pp. 282-288.
Durelli, A. J., Obert, L., and Parks, V. J., 1968, "Stress
Required to Initiate Core Discing," Trans. SME-AZME,
Vol. 241, No. 3, Sept., pp. 269-276.
Duvall, W. I., 1976, "General Principles of Underground
Opening Design in Competent Rock," Monograph on Rock
Mechanics Applications in Mining, W. S. Brown, S. J.
Green, W. A. Hustrulid, eds., AIME, New York, pp.
3 1-39.
Ellis, W. L.. and Ege, J. R., 1976, "Determination of In
Situ Stress in U12G Tunnel, Rainier Mesa, Nevada Test
Site, Nevada," 474-219, US Geological Survey, Jan.
Habid, P., and Marchand, R., 1952, "Mesures des Pressions
de Terrains par L'Essai de Veriu Plat," Annales de
I'Znstitut Technique drc Batiment et des Travaux Publics,
Serie Sols et Foundations, NO. 58.
Haimson, B. C., 1974, "Determination of Stresses in Deep
Holes and Around Tunnels by Hydraulic Fracturing,"
Proceedings, Rapid Excavation and Tunneling Conference,
Vol. 2, AIME, New York, pp. 1539-1560.
Haimson, B. C., et al., 1974, "Deep Stress Measurement in
Tuff at the Nevada Test Site," Proceedings, Third Congress
of the International Society for Rock Mechanics, Denver.
Hawkes, I., 1968, "Theory of the Photoelastic Biaxial Strain
Gage," International Journal of Rock Mechanics and
Mining Science, Vol. 5, NO. 1, pp. 57-64.

Hawkes, I., and Hooker, V. E., 1974, "The Vibrating Wire


Stress Meter," Proceedings, Third Congress of the International Society for Rock Mechanics, Denver.
Hawkes, I., and Moxon, S., 1965, "The Measurement of In
Situ Rock Stress Using the Photoelastic Biaxial Gage with
Core-Relief Technique," International Journal of Rock
Mechanics and Mining Science, Vol. 2, pp. 405-419.
Hedley, D. G., and Wilson, J. C., 1975, "Rock Mechanics
Applications in Canadian Underground Mines," CZM
Bulletin, Nov.
Hiramatsu, Y . , Niwa, Y., and Oka, Y., 1957, "Measurement
of Stress in Field by Application of Photeoelasticity,"
Technical Report, Vol. 7, No. 37, Engineering Research
Institute, Kyoto University.
Hooker, V. E., Aggson, J. R., and Bickel, D. L., 1971, "Report on In Situ Determination of Stress, Rainier Mesa,
Nevada Test Site," Internal Report, US Bureau of
Mines.
Hooker, V. E., Aggson, J. R., and Bickel, D. L., 1974, "Improvements in the Three-Component Borehole Deformation Gage and Overcoring Techniques," Report of Investigations 7894, US Bureau of Mines.
Hooker, V. E., Bickel, D. L., and Aggson, J. R., 1972, "In
Situ Determination of Stress in Mountainous Topography,"
Report of Investigations 7654, US Bureau of Mines.
Hooker, V. E., and Bickel, D. L., 1974, "Overcoring Experiments and Techniques Used in Rock Stress Determination," Information Circular 8618, US Bureeu of Mines.
Hooker, V. E., and Duvall, W. I., 1966, "Stresses in Rock
Outcrops near Atlanta, GA," Report of Investigations
6860, US Bureau of Mines.
Hooker, V. E., and Duvall, W. I., 1971, "In Situ Rock
Temperature Stress Investigations in Rock Quarries," Report of Investigations 7589, US Bureau of Mines.
Hooker, V. E., and Johnson, C. F., 1969, "Near-Surface
Horizontal Stresses Including the Effects of Rock Anisotrophy," Report of Investigations 7224, US Bureau of
Mines.
Hooker, V. E., Nicholls, H. P., and Duvall, W. I., 1964,
"In Situ Stress Determinations in a Lithonia Gneiss Outcrop," Proceedings, Annual Meeting, Eastern Section,
Seismological Society of America, Ann Arbor, Oct.
Horino, F. G., and Aggson, J. R., 1976, "Pillar Failure
Analysis and In Situ Stress Determination at the Fletcher
Mine, Bon Terre, Missouri," Report of Investigations US
Bureau of Mines.
Hubbert, M. K., and Willis, D. G., 1957, "Mechanics of
Hydraulic Fracturing," Transactions, American Society of
Mechanical Engineers, Vol. 210, pp. 153-168.
Kehle, R. O., 1964, "The Determination of Tectonic Stresses
Through Analysis of Hydraulic Well Fracturing," Journal
of Geophysical Research, Vol. 69, NO. 2, pp. 259-273.
Leeman, E. R., 1964, "A Trepanning Stress Relieving Factor
for Rock Stress Measurements," Proceedings, Sixth Symposium on Rock Mechanics, University of Missouri, Rolla,
pp. 407-426.
Leeman, E. R., 1969, "The Doorstopper and Triaxial Rock
Stress Measuring Instruments Developed by CSIR,"
Journal of the South African Znstitute of Mining and
Metallurgy, Vol. 69, NO. 7, pp. 305-339.
Lieurance, R. S., 1933, "Stress in Foundations at Boulder
Dam." Technical Memo 346, US Bureau of Reclamation,

co.

Merrill, R. H., 1970, "The Factors that Influence the Design on Slope Walls," Proceedings, Sixth International
Mining Congress, Madrid, Preprint, 1-C, 2.
Merrill, R. H., and Peterson, J. R., 1961, "Deformation of a
Borehole in Rock," Report of Investigations 5881, US
Bureau of Mines.
Merrill, R. H., et al., 1964, "Stress Determination by Flat
Jack and Borehole Deformation Methods," Report of Investigations 6400, US Bureau of Mines.

DESIGN CONTROL
Obert, L. A., 1962, "In Situ Determination of Stress in
Rock," Mining Engineering, Vol. 14, Aug., pp. 51-58.
Obert, L. A., 1972, "Philosophy of Design, Proceedings of a
Bureau of Mines Technical Transfer Seminar, Denver,
Colorado," Information Circular 8585, US Bureau of
Mines.
Obert, L. A., 1964, "Triaxial Method for Determining the
Elastics Constants of Stress Relief Cores," Report of Investigations 6490, US Bureau of Mines.
Obert, L. A., and Duvall, W. I., 1967, Rock Mechanics and
the Design of Structures in Rock, John Wiley and Sons,
New York, pp. 245, 419, and Chapter 16.
Obert, L. A., Merrill, R. H., and Morgan, T. A., 1962,
"Borehole Deformation Gage for Determining the Stress in
Mine Rock," Report of Investigations 5978, US Bureau
of Mines.
Obert, L. A., and Stephenson, D. E., 1965, "Stress Conditions Under Which Core Discing
- Occurs," Trans. SMEAIME, Vol. 23 1, Sept.
Palmer. J. H. L.. and Lo. K. Y.. 1976. "In Situ Stress
~easurementsih Some Near-surface ~ 6 c kFormationsThorold, Ontario," Canadian Geotechnical Review, Vol.
13, No. 1.
Panek, L. A., 1966, "Calculation of the Average Ground
Stress Components from Measurements of the Diametral
Deformation of a Drill Hole," Report of Investigations
6732, US Bureau of Mines.
Panek, L. A., and Stock, J. A., 1964, "Development of a
Rock Stress Monitoring Station Based on the H a t Slot
Method of Measuring Existing Rock Stress," Report of
Investigations 6537, US Bureau of Mines.
Raleigh, C. B., Healy, J. H., and Bredehoeft, J. D., 1972,
"Faulting and Crustal Stress at Rangely, Colorado, in
Flow and Fracture of Rocks," Geophysical Monograph 16,
American Geophysical Union.

Roberts, A., et al., 1964, "A Laboratory Study of the Photoelastic Stress Meter," International Journal o f Rock Mechanics and Mining Science, Vol. 1, No. 3, May.
Rocha, M., Lopes, J. B., and Da Silva, J. N., 1966, "A New
Technique for Applying the Method of the Flat Jack in
Determination of Stresses Inside Rock Masses," Proceedings, First Congress of the International Society for Rock
Mechanics, Lisbon.
Scheidegger, A. E., 1960, "On the Connection Between
Tectonic Stresses and Well Fracturing Data," Pure and
Applied Geophysics, Vol. 46, No. 2, pp. 66-76.
Skinner, E. H., Waddell, G. G., and Conway, J. P., 1974,
"In Situ Determination of Rock Behavior by Overcore
Stress Relief Method, Physical Property Measurements,
and Initial Deformation Method," Report of Investigations
7962, US Bureau of Mines.
Smith, R. L., 1972, "Mine Installation of Two Bureau of
Mines Hydraulic Pressure Cells and a Borehole Deformation Gage, Proceedings of a Bureau of Mines Technical
Transfer Seminar, Denver, Colorado," Information Circular 8585, US Bureau of Mines.
Van Heerden, W. L., 1976, "Practical Application of the
CSIR Triaxial Strain Cell for Rock Stress Measurements,"
Proceedings, Investigation of Stress in Rock-Advances in
Stress Measurement, Sydney, Aug., pp. 1-6.
Van Heerden, W. L., and Grant, F., 1967, "A Comparison
of Two Methods for Measuring Stress in Rock," Znternational Journal o f Rock Mechanics and Mining Science,
Vol. 4, pp. 367-382.
Worotnicki, G., and Walton, R. J., 1976, "Triaxial Hollow
Inclusion Gauges for the Determination of Rock Stresses
In Situ," ISRM Symposium on Advances in Rock Stress
Measurement, Sydney, Australia.

Вам также может понравиться