Fulltext01 PDF
Fulltext01 PDF
Fulltext01 PDF
TECHNICAL REPORT
Seismicity in Mines
A Review
Kristina Larsson
This literature review is part of a joint research project between SveBeFo, Boliden Mineral
AB, LKAB, and Luleå University of Technology. The project is aimed at increasing the
knowledge of seismicity induced by mining in Sweden. The financial support for the project
is provided by SveBeFo, Boliden Mineral AB and LKAB.
The work presented in this report is the result of a literature study and several discussion
meetings with the project reference group. The reference group consists of M.Sc. Christer
Andersson, SKB (Swedish Nuclear Waste Management), Mr. Tomas Franzén, Research
Director at SveBeFo, Tech. Lic. Lars Malmgren, LKAB, M.Sc. Per-Ivar Marklund, Boliden
Mineral AB, Professor Arne Myrvang, NTNU, Dr. Jonny Sjöberg, SwedPower AB, and my
supervisor Professor Erling Nordlund, Division of Rock Mechanics at Ltu.
Kristina Larsson
i
ii
SUMMARY
The virgin stress state in the rock mass depends on several factors, for instance gravity and
tectonics. Around a mine the stress state is disturbed, which leads to locally increased or
decreased stresses. The rock mass in Sweden is generally composed of high strength brittle
rock types, so the risk of violent failures increases with increasing depth of mining due to
increasing stress levels. The problem with violent failures is that they are a safety risk for
mine personnel. During the past few years an increased number of violent failures causing
fallouts have been observed in both LKAB and Boliden mines. These types of failures are
generally called rockbursts, but different mines give different meaning to the phenomena.
Therefore the terms seismicity and seismic event will be introduced here. Seismicity is the
rock mass response to deformation and failure. A seismic event is the sudden release of
potential or stored energy in the rock. The released energy is then radiated as seismic waves.
A rockburst is defined as a mining-induced seismic event that causes damage to openings in
the rock.
In this literature report, mining induced seismicity is reviewed and the influencing factors
described. The state of stress is one important factor that influences the occurrence of
rockbursts; hence the state of stress in Scandinavia is described briefly, and examples of how
the most common Swedish mining methods influence the stress field are given. The behavior
of the rock mass in different situations is described, dealing with the relationship between
stress state and seismicity and rock response to energy changes during excavation. Different
rockburst mechanisms are described including failure mechanisms and type of damage
caused. Rockburst classifications and the varying definitions of rockbursts are briefly
summarized. Seismic monitoring is an important tool for many mines today, and is therefore
reviewed including a description of the most common seismological terms that are used to
describe seismicity. Some different methods of predicting and explaining rockbursts are
described. Methods of preventing seismicity, for example different ways of destressing the
rock mass and reinforcement have also been reviewed.
iii
iv
SAMMANFATTNING
Det primära spänningstillståndet i bergmassan beror av flera faktorer, bland annat gravitation
och tektonik. Spänningstillståndet runt en gruva är stört, vilket leder till lokalt ökade eller
minskade spänningar. Bergmassan i Sverige består generellt av höghållfasta spröda bergarter,
varför risken för våldsamma brottförlopp ökar med ökande brytningsdjup på grund av de
högre spänningarna. Problemet med de våldsamma brotten är att de utgör en säkerhetsrisk för
personalen. Under de senaste åren har antalet våldsamma brott som orsakat utfall ökat i både
Bolidens och LKABs gruvor. Dessa typer av brott kallas generellt för smällberg (eng.
rockburst), men olika gruvor har olika uppfattning om vad smällberg är. Därför introduceras i
denna rapport begreppet seismicitet och seismisk händelse. Seismicitet är bergmassans
respons på deformation och brott. En seismisk händelse är en plötslig frigörelse av potentiell
eller lagrad energi i berget, vilken avges i form av seismiska vågor. Smällberg definieras som
en seismisk händelse som orsakar en skada (utfall) på öppningar i bergmassan.
v
vi
LIST OF SYMBOLS AND ABBREVIATIONS
vii
Wk = kinetic or seismic energy
Wr = released energy
Ws = energy absorbed in support deformation
Wt = change in potential energy
Uc = stored strain energy
Um = stored energy in removed rock
Us = energy stored in spring
Uu = increase in stored energy
U1 = initial strain energy of test cell in MGW
U2 = strain energy after event in MGW
K = stiffness of spring or surrounding rock
F = force, N
N = normal force acting perpendicular to surface of interest
µs = static coefficient of friction
µd = dynamic coefficient of friction
A = area
s = speed
φ = friction angle
cs = cohesion
α = strike angle, or azimuth
β = dip angle
λ = rake
D = average shear displacement
d = distance
Λ0 = amplitude
Λ = maximum amplitude
∆ = focal depth
T = period of the measured wave
h = range
ϕ = epicentral distance, °
Ξ = epicentral distance in kilometers
Γ = instrument magnification factor
∆σ = stress drop
viii
r0 = radius of fault
f0 = corner frequency
ML = local or Richter magnitude
M0 = seismic moment
MS = surface wave magnitude
mb = body wave magnitude
Mw = moment magnitude
Mn = Nuttli magnitude
ES, EP = seismic energy in S-, and P-wave respectively
Et = total seismic energy
N = number of events
a = constant
b = relation between small and large events in a certain time interval
ERR = Energy Release Rate
ESS = Excess Shear Stress
VESS = Volume Excess Shear Stress
MGW = Modeled Ground Work
LERD = Local Energy Release Density
ID = departure index
Pi , Pj = parameters
RHI = rockburst hazard indication
Ψ = function with value 1 or 0 depending on the departure index
Q1 = stress parameter in cell evaluation method
Q2 = strain energy parameter in cell evaluation method
ST = factor of safety
tn, tn’ = traction, before and after event
un, un’ = displacement, before and after event
ix
x
TABLE OF CONTENTS
Preface......................................................................................................................................... i
Summary ...................................................................................................................................iii
Sammanfattning ......................................................................................................................... v
List of symbols and abbreviations............................................................................................vii
Table of contents ....................................................................................................................... xi
1 Introduction ........................................................................................................................ 1
1.1 Background ................................................................................................................ 1
1.2 Objective and outline of report................................................................................... 2
2 Rock stress.......................................................................................................................... 3
2.1 State of stress in the rock mass................................................................................... 3
2.2 Classification of rock stresses .................................................................................... 3
2.3 Virgin stresses or in-situ stresses ............................................................................... 4
2.3.1 Gravitational stresses.......................................................................................... 4
2.3.2 Tectonic stresses................................................................................................. 8
2.3.3 Residual stresses................................................................................................. 9
2.4 Stresses in Scandinavia ............................................................................................ 10
2.5 Secondary stresses or induced stresses..................................................................... 13
2.5.1 General ............................................................................................................. 13
2.5.2 Mining induced stresses ................................................................................... 14
3 Rock mass behavior ......................................................................................................... 19
3.1 Response of rock to energy changes ........................................................................ 19
3.2 Compressive stress and seismicity ........................................................................... 23
3.3 Shear stress and seismicity....................................................................................... 25
4 Rockburst source mechanisms ......................................................................................... 29
4.1 Seismic events associated with stopes ..................................................................... 29
4.1.1 Strain burst ....................................................................................................... 29
4.1.2 Pillar burst ........................................................................................................ 29
4.2 Seismic events associated with geologic discontinuities ......................................... 30
4.2.1 Fault slip........................................................................................................... 30
4.2.2 Shear rupture .................................................................................................... 31
4.3 Rockburst definitions ............................................................................................... 31
4.3.1 Definition ......................................................................................................... 31
xi
4.3.2 Classification / categorization of rockbursts .................................................... 34
5 Seismic monitoring and earthquake seismology.............................................................. 39
5.1 Introduction .............................................................................................................. 39
5.2 Seismic monitoring .................................................................................................. 39
5.3 Earthquake seismology ............................................................................................ 41
5.3.1 Faults and seismic waves ................................................................................. 42
5.3.2 Seismic moment and the moment tensor.......................................................... 45
5.3.3 Earthquake focal mechanisms and radiation patterns ...................................... 49
5.3.4 Earthquake location.......................................................................................... 50
5.3.5 Stress drop ........................................................................................................ 52
5.3.6 Earthquake magnitude...................................................................................... 53
5.3.7 Seismic energy ................................................................................................. 57
5.3.8 Corner frequency.............................................................................................. 58
5.3.9 Intensity scale................................................................................................... 59
6 Description of seismicity.................................................................................................. 63
6.1 Introduction .............................................................................................................. 63
6.2 Describing seismicity without seismic records ........................................................ 63
6.2.1 Energy Release Rate (ERR) ............................................................................. 63
6.2.2 Excess Shear Stress (ESS)................................................................................ 65
6.2.3 Volume Excess Shear Stress (VESS)............................................................... 67
6.3 Describing seismicity using seismic records............................................................ 68
6.3.1 Departure indexing method.............................................................................. 68
6.3.2 Cell evaluation method - Modeled Ground Work (MGW).............................. 70
6.3.3 Local Energy Release Density (LERD) ........................................................... 73
6.3.4 Seismic Hazard Scale (SHS) ............................................................................ 74
7 Prevention and control of seismicity................................................................................ 77
7.1 Preconditioning and destressing............................................................................... 77
7.1.1 South Africa ..................................................................................................... 78
7.1.2 Sweden ............................................................................................................. 81
7.2 Fluid injection .......................................................................................................... 88
7.3 Reinforcement .......................................................................................................... 89
7.3.1 Introduction ...................................................................................................... 89
7.3.2 Support functions ............................................................................................. 89
7.3.3 Support capacity............................................................................................... 90
xii
7.3.4 Support systems................................................................................................ 92
8 Concluding remarks ......................................................................................................... 95
9 References ........................................................................................................................ 97
xiii
xiv
1 INTRODUCTION
1.1 Background
The virgin state of stress in the Earth’s crust depends on, among other factors, gravity and
tectonics. In, or near, a mine the stress state is disturbed by the mining activities leading to
locally increased or decreased stresses. The extent and the consequences of a failure are
controlled by three fundamental elements: geomechanical conditions, stress state and mining
activities. The term geomechanical conditions include intact rock with geological structures
(joints, faults, folds etc.) and it is characterized by the strength and stiffness of the intact rock
and the joints, water pressure etc. For failure to occur the interaction of all these three factors
is required. Depending on the characteristics of the factors, different failure modes can occur.
When the rock mass consists of low strength rock types and also has a low quality, high stress
levels may lead to massive failures in the rock mass and large plastic deformations. If the rock
mass is of high quality and consists of brittle high strength rock types, the risk for violent
failures increases. The failures can, in such cases, be relatively restricted to the area closest to
the opening or result in rapid movements along geological structures leading to earthquake-
like phenomena.
The increasing mining depth in the Swedish underground mines leads to increasing stress
levels and thus to increased risk of instability in drifts, ramps, shafts and other underground
structures. The high quality of the rock mass in the areas mined today, in combination with
increasing stresses will most likely lead to violent failures with an increasing extent and an
increase in the volume of fallouts in the future. At larger depths the risk of other phenomena
like activation of faults also increases. Slip on a fault releases large amounts of energy that
can damage underground constructions. The problem with violent failures is that they are a
safety risk for mine personnel. People working close to the face are very vulnerable if the face
or roof starts to eject material. Popping noises sometimes precede these kinds of failure, but
not always. Popping may also occur without ejection of material, so it may be difficult to
determine the risk.
During the past few years an increased number of violent failures causing fallouts have been
observed in both LKAB and Boliden mines in Sweden. Despite apparently similar conditions
the problems with brittle and violent failures have been observed to vary in intensity. This
1
project was started to appraise the extent of the problem and to find the mechanisms behind
the violent failures.
These types of failures are generally called rockbursts. The term is used without
discrimination, so a failure that is called a rockburst in one mine may not be considered so in
another. Therefore the terms seismicity and seismic event will be introduced here. Seismicity
is the rock mass response to deformation and failure. A seismic event is the sudden release of
potential or stored energy in the rock. The released energy is then radiated as seismic waves.
A rockburst is defined as a mining-induced seismic event that causes damage to openings in
the rock. These definitions have been used by for instance Cook (1976), Salamon (1983) and
Ortlepp and Stacey (1994). The vagueness of the term rockburst is evident here, since this
definition says nothing of how large the damage should be. An event displacing 1 kg of rock
and one that closes an entire drift are both rockbursts.
Chapter 2 of this report describes the stress state in the rock mass, both virgin and induced
stresses. The state of stress is one important factor that influences the occurrence of
rockbursts. The project is about problems caused by high stress levels, so the state of stress in
Scandinavia is described briefly. Examples of how mining influences the stress field are
given. Chapter 3 describes rock mass behavior in different situations, dealing with the
relationship between stress state and seismicity and rock response to energy changes during
excavation. In Chapter 4, the different rockburst mechanisms are described including failure
mechanisms and type of damage caused. Rockburst classifications and the varying definitions
of rockbursts are briefly summarized. Chapter 5 is a description of seismic monitoring in
general, and a summary of the most common seismological terms that are used to describe
seismicity. Chapter 6 describes some different methods of explaining and predicting
rockbursts. Chapter 7 deals with methods of preventing seismicity, for example different ways
of destressing the rock mass and reinforcement. Chapter 8 gives conclusions drawn from the
literature review regarding influencing factors.
2
2 ROCK STRESS
Rock stresses
3
2.3 Virgin stresses or in-situ stresses
Virgin stresses are the stresses that exist in the rock mass before any disturbance (e.g.,
excavation) has occurred. The virgin stresses comprise gravitational, tectonic, residual, and
terrestrial stresses. Terrestrial stresses are the stresses induced by diurnal and seasonal
variations of temperature, Moon pull and the Coriolis force (Amadei and Stephansson, 1997).
These are often neglected but can affect stress measurements at shallow to very shallow
depths. They have very little influence on the stress state at depth in Scandinavia and will be
left out of the following description.
where ρ is the density of the rock mass (kg/m3), g is the gravity acceleration (9.81 m/s2), and z
is the depth below ground surface.
The horizontal component due to gravitational loads depends on the rock mass properties. If
the material can be considered linear elastic and isotropic and a one-dimensional state of
strain is assumed, the average horizontal stress is defined, by for instance Herget (1988), as
ν
σH = σv Eq. 2-2
1 −ν
where ν is Poisson’s ratio, which may vary between 0.15 and 0.35 for most rock types.
If the rock material cannot be considered elastic, and the rock cannot resist shear stresses in
the long term (i.e., creep due to viscosity will occur) then after some time the magnitude of
the horizontal component will equal the vertical component. This is called a hydrostatic or
lithostatic stress field (Herget, 1988). In a lithostatic stress field the weight of the overburden
acts in all directions. In short, the horizontal component due to gravitational factors can be
expressed as
σ H = kσ v Eq. 2-3
4
where k is a factor that varies from 0 to 1 depending on the restraints on displacement.
So far we have assumed that the rock mass is isotropic, but this is seldom true on a larger
scale. Anisotropy is common in rock and can be caused by for example bedding,
stratification, schistosity planes or jointing. Foliated metamorphic rock types (schists,
gneisses), stratified sedimentary rocks (sandstone, limestone) and rocks with regular, closely
spaced joint sets have anisotropy as a general characteristic. Amadei et al. (1987) studied the
stress field induced in an anisotropic rock mass under gravitational loading assuming a one-
dimensional state of strain. They assumed that the rock mass was homogeneous, linearly
elastic and that the material was either orthotropic or transversely isotropic. Transversely
isotropic means that the rock mass has different properties in two directions, i.e., there are
isotropic planes, see Figure 2.2a. Orthotropy means that the rock mass has different properties
in all three directions, i.e.,, there are three orthogonal planes of elastic symmetry, see Figure
2.2b.
σz
a) σz b)
σx = σy σx
σy = σx σy
For an orthotropic rock mass with only gravitational loading, no lateral displacements and a
horizontal ground surface, Amadei et al. (1987) derived expressions for the two horizontal
stresses. They found that gravitational loading of an orthotropic rock mass induces unequal
principal stresses in the horizontal plane. They also suggest that σx and σy are strongly
dependent on the degree of anisotropy. For some ranges of the anisotropic rock properties, the
horizontal stresses can exceed the vertical stresses, which is not possible in the isotropic
solution. In this study by Amadei et al. (1987) it was assumed that the strata were either
parallel or perpendicular to the horizontal ground surface. In a study by Amadei and Pan
5
(1992) they examined the gravity-induced stresses in orthotropic and transversely isotropic
rock with strata inclined with respect to a horizontal ground surface, see Figure 2.3.
y β
x
z
P
Figure 2.3. Model of transversely isotropic rock mass with strata inclined at a dip, β.
From this model they concluded that if one of the planes of symmetry is parallel to the y-axis
the stress components σx, σy and σz are principal stresses. σx and σy are horizontal stresses,
and σz is the vertical stress. The conclusion that can be drawn from both these studies is that
for orthotropic and transversely isotropic rock masses, the horizontal stresses parallel to the
strike and dip direction of the strata are always principal stresses and are not equal. The stress
field induced by gravity is three-dimensional for anisotropic rock masses and the horizontal
stresses can be smaller than, equal to, or larger than, the vertical stress.
Stratification is common in both volcanic and sedimentary rocks. Lithology and relative
stiffness between different layers can cause the in-situ stresses to vary considerably between
different layers. Warpinski and Teufel (1991) made stress measurements (using hydraulic
fracturing) at Rainier Mesa, Nevada. They found that the minor principal stress could vary
between layers and that the stress change could occur on a scale of about 1 m, if the difference
in material properties was large enough. From these studies they concluded that the horizontal
stresses at Rainier Mesa depended on material properties of the various layers, topography
and bedding. They observed several induced hydraulic fractures that changed direction or
stopped at a fault, and also measured stress magnitudes that could vary by as much as 2 MPa
and have a change of orientation of 20°- 40°. The stress measured was the minimum in-situ
stress, and the average was for one hole 3.92 MPa, and for the other 4.21 MPa, which means a
50 % variation in stress across the fault.
6
The distribution and magnitude of horizontal stresses may also be affected by variations in
rock mass geology and the existence of geological structures and other heterogeneities, for
example orebodies. Local changes in the in-situ stress field can be expected when crossing a
persistent discontinuity. Hudson and Cooling (1988) identified three different cases due to the
relative stiffnesses of the material in the discontinuity versus the surrounding rock material. If
the discontinuity is open, the major principal stress will change orientation to become parallel
to the discontinuity, while the minor principal stress will become perpendicular to the joint
and be reduced to zero. If the joint filling has roughly the same stiffness as the surrounding
rock, the stresses may be virtually unaffected. If the discontinuity filling is stiffer than the
surrounding rock, the major principal stress changes orientation to become perpendicular to
the discontinuity while the minor principal stress becomes parallel.
According to Amadei and Stephansson (1997) research has shown that the minor and major
horizontal stresses (in Australia) often align with orebodies. An orebody can be seen as a
heterogeneity, and the disturbance it causes in the in-situ stress field can be understood using
the analogy of a solid inclusion embedded into an infinite medium. Using the theory of
elasticity it can be shown that the stresses in a solid inclusion embedded in a continuous
medium differ from the stresses in the surrounding material. If the surrounding material is of
infinite dimensions (isotropic or anisotropic) the stresses in the inclusion are uniform, but if
the surrounding material is finite the stresses in the inclusion vary from point to point. The
inclusion also affects the surrounding material. How large the zone of influence is depends on
the difference in stiffness. In practice this means that heterogeneities can cause disturbances
(stress concentrations) in the stress field large enough to cause rockbursts or other stability
problems during excavation.
A very common assumption in rock mechanics is that the principal stresses are vertical and
horizontal. This is not always true, especially at shallow depths when the ground surface is
not horizontal. If we look at a very rocky area consisting of high peaks and deep valleys and
only consider gravity-induced stresses and allow no horizontal displacements, we will find
that the principal stresses are parallel and normal to the ground surface, see Figure 2.4. As the
depth increases, the effect of the rugged topology is reduced and the principal stresses resume
the same orientations that they would have had if the ground surface was horizontal.
7
Figure 2.4. Effect of topography.
5 3 4
7
1
6
Figure 2.5. Forces causing tectonic stresses (after Amadei and Stephansson, 1997).
One effect of tectonics (ridge push) is that the stresses can be very much higher than the depth
would indicate, which can lead to problems related to high stresses even at shallow depths.
For instance in Norway during construction of the Kobbskaret road tunnel (Myrvang et al.,
8
1997) heavy spalling occurred with overburdens of only a few tens of meters. The tunnel
crosses a mountain range and the measured major principal stress (σ1 = 27 MPa) was oriented
parallel with the strike of the mountain range and thus normal to the tunnel axis. The
overburden varied from 25 to 600 m and spalling occurred along almost the whole tunnel
length. Pushing from the Midatlantic ridge is the primary reason that the horizontal stresses in
Scandinavia are significantly higher than the vertical stresses.
These requirements can be explained by studying a granite block that was formed under high
temperature and high pressure. At the time of crystallization all constituent grains interlock
perfectly. As the block cools the grains will change shape, the amount of change depending
on their respective elastic moduli, coefficient of thermal expansion, and non-preferred
orientation of anisotropic grains. If all the grains could strain freely and contained no
heterogeneities, each would assume a stress free state. This is not possible since the
compatibility of the grain contacts and integrity of the block would not be maintained. Due to
the confining effect of neighboring grains, stresses are induced, which are called the residual
stresses.
The importance of residual stresses is difficult to estimate, but they are believed to be at least
partially responsible for phenomena such as rockbursts, surface spalling and sheet jointing.
The residual stresses also contribute to both instantaneous and time-dependent movement
9
when rock is removed – both on a small scale (drilling and coring) and on a large scale (whole
excavations) (Amadei and Stephansson, 1997).
- lateral variation of plate boundary forces, caused by the Jan Mayen Fracture Zone,
- a locally radial ridge push, caused by the Iceland hot spot, that is superimposed on the
lateral plate boundary forces, and
- the physical properties of the Fennoscandian lithosphere, which reduce the mean stress level
of the lithosphere and allow local inhomogeneities (density, strength) to have an important
influence on the stress field.
The conclusion is that the stress field in Scandinavia is a combination of plate boundary
forces in combination with local sources, for example flexural stresses. The large scatter may
also suggest that the horizontal principal stresses are very close in magnitude, an observation
that is supported by the occurrence of focal mechanisms of different styles close to each other.
Stephansson (1993) compiled data from stress measurements (by hydraulic fracturing) in
Scandinavia and found that the major and minor principal horizontal stresses varied with
depth, z, as:
10
Stephansson (1993) also summarized the results from all overcoring stress measurements
made in Sweden, and found that the major, intermediate and minor principal stresses varied
with depth, z, as (z < 1000 m):
These data are taken from the Fennoscandian Rock Stress Data Base (FRDSB), which forms a
subset of the World Stress Map Project. This is a database of tectonic stresses in the Earth’s
crust, where the tectonic stress directions are determined by using four categories of stress
indicators (Reinecker et al., 2004):
The database contains more than 13000 data points that are ranked according to their
reliability as a tectonic stress indicator. There are five categories, A-E, where A is the most
reliable. Normally only data from categories A-C are displayed on the maps. For Scandinavia,
the most recent map is shown in Figure 2.6. From the map it can be seen that the directions of
the major horizontal stress vary to a large extent.
11
0˚ 10˚ 20˚ 30˚
Method:
focal mechanism
breakouts
drill. induced frac.
borehole slotter
overcoring
hydro. fractures
geol. indicators
Regime:
70˚ NF SS TF U 70˚
Quality:
A
B
C
(2004) World Stress Map
60˚ 60˚
12
2.5 Secondary stresses or induced stresses
2.5.1 General
Secondary stresses (or induced stresses) are the result of redistribution of the primary stresses
because of a disturbance. The disturbance can be either natural, like a change in conditions
(drying, swelling, or consolidation), or be caused by human actions (excavation, pumping, or
energy extraction) (Herget, 1988). If the rock is assumed to be elastic and under uniaxial
loading, the redistributed stresses form “streamlines” around the opening, which in fact are
major principal stress trajectories. An example for a circular excavation is shown in Figure
2.7.
Figure 2.7. Major principal stress trajectories around a circular cross section.
Shorter distance between the flow lines indicates an increase in stress, and a wider spacing
indicates a decrease in stress compared to the virgin state of stress. At a distance of about
three diameters away from the boundary of the opening the stresses are virtually undisturbed;
hence a single excavation only disturbs the stress field very locally
When the cross section is changed from circular to some other shape, the difficulty in finding
closed form analytical solutions for the state of stress around the excavation increases
dramatically. A rounded shape gives a smoother flow around the excavation, compared to
e.g., a square or rectangular shape. The corners become points of stress concentration. If the
cross-section of the opening is much longer in one direction, and has sharp corners, zones
with decreased stresses may form in the middle of the roof, see Figure 2.8, while there are
zones of increased stresses in the walls.
13
Figure 2.8. Flow lines around parallel tunnels with rectangular cross section (Hoek and
Brown, 1982).
If more excavations are added close to the existing one in Figure 2.7, the stress field will be
disturbed over a larger area. The stress distribution around an opening will influence the state
of stress around the others, making it difficult to calculate the secondary stresses. The
simplest case is that of two parallel circular tunnels. The stress field around the tunnels is still
not too complicated and can be treated analytically. When the cross-sections become more
complicated (i.e., not circular), or the number of excavations exceeds two, numerical methods
have to be used to study how the openings will influence each other, and what the resulting
secondary stresses are.
14
the orebody. The hangingwall fractures and collapses – thus, the ground surface must be
allowed to subside. Only a negligible part of the stresses pass through the caved rock because
of its low stiffness compared to the rest of the rock mass. The major part of the stress
perpendicular to the orebody is redistributed under the caved zone, see Figure 2.9.
Broken ore
Footwall drifts
Figure 2.9. Schematic figure of stress redistribution under the caved zone.
The deeper the mine the higher the stress concentrations in the bottom become. The stress
perpendicular to the orebody is highest about 50-100 m below the bottom of the cave (Sjöberg
et al., 2001) and drifts oriented parallel to the orebody (e.g., footwall drifts) become highly
stressed. The stress parallel to the orebody is also partly redistributed under the cave, but in
this case the major part is redistributed “horizontally” to the foot- and hangingwall, see Figure
2.10. This stress redistribution affects crosscuts and access drifts perpendicular to the
orebody.
Hangingwall
Principal stress
trajectories
Orebody
Footwall
15
As an example of the stress redistribution around the cave we study a footwall drift on the
development level. When it is excavated it is subjected to a stress field that is close to the
virgin stress field in the area, both regarding orientation and magnitude of the major principal
stress, see Figure 2.11a. As mining and the caving progresses downward to the level of the
drift, the secondary stresses around the footwall drift increase as the horizontal stresses are
forced under the cave. The direction of the major principal stress (i.e., the stress that is not
disturbed by the opening itself, but from the caved area) rotates so that it is not horizontal
anymore, see Figure 2.11b. The stress level around the drift can be high enough to cause
compressive failure at the boundary. When the extraction level has passed the level of the
footwall drift by 100 m or more, the stresses decrease to a level that can be lower than the
virgin stress state, and the major principal stress has rotated to become more or less parallel to
the orebody, see Figure 2.11c. At this time fallouts of broken rock are likely to occur, since
the stresses are too low to retain the blocks.
σ1
a) b) c)
σ1
σ1 σ1
σ1
σ1
Figure 2.11. Rotation of stress field during the life of a footwall drift.
Permanent installations like ore passes and ventilation shafts are also affected by the stress
redistribution. All these excavations are located some distance away from the orebody and are
constructed years before the ore extraction reaches that level. A shaft for instance is subjected
to varying stress conditions along its length. The bottom part may be subjected to an
undisturbed stress field, the middle part of the shaft may be very highly stressed, while the top
part has a significantly lower stress state (almost destressed). This is caused by the varying
stress field with respect to magnitude and orientation.
16
lift consists of eight to ten slices, giving a lift height of 40 to 60 m, and then a sill pillar is left.
As each slice is mined out, it is backfilled and serves as a working platform for the next lift.
The stress situation of the overhand cut-and-fill method can be described as in Figure 2.12. As
mining progresses upward, the loading of the orebody above and below the stope increases.
When the stope approaches the mined out (and backfilled) excavation above, the remaining
ore forms a horizontal sill pillar subjected to very high stresses. If the stresses in the pillar
approach the strength, failure will occur either in the pillar itself or in the sidewall rock.
Figure 2.12. Stress increase in stope roof as mining progresses upward, after Krauland
and Söder (1989).
When the sill pillar has failed, the decrease in stiffness leads to redistribution of stresses over
and under the stope. Sometimes uppers are used to remove all, or almost all, of the sill pillar.
Often the mine started out as an open pit and then went underground leaving a crown pillar,
which will be increasingly fractured as mining takes place underneath, similar to the sill pillar
described above. When it has failed completely, the stress redistribution around the mine is
similar to that which occurs in sublevel caving.
17
18
3 ROCK MASS BEHAVIOR
a) Case I b) Case II
tn-1 tn-1
tn tn
t1 t1
V tm+1
V
tm+1
t2 t2
tm tm Vm Sm
t3 t3
tm-1 tm-1
Si t4 Si t4
S S
The following analysis assumes that the rock mass is linearly elastic and that no energy is
consumed in fracturing or non-elastic deformation. The stresses that acted on the removed
rock are transferred to the surrounding rock mass, which increases its stored strain energy
(Uc). If the excavations are supported then some energy is absorbed in deforming the support
19
(Ws). The energy remaining is referred to as released energy (Wr). The law of conservation of
energy gives
Wt + U m = U c + Ws + Wr Eq. 3-1
If the rock was removed instantaneously this would cause vibrations in the rock mass, and
equilibrium would be restored by damping and in the process seismic energy or kinetic energy
(Wk) would be dissipated (Hedley, 1992). The energy that has to be released is the sum of the
energy stored in the removed rock (Um) and the kinetic energy (Wk). For elastic conditions
there are no more alternatives, which means that
W r = U m + Wk Eq. 3-2
The seismic energy component contributes to the damage caused by a rockburst, and it is
seismic energy that is recorded by seismic systems. Combining Equations 3-1 and 3-2 gives
Wk = Wt − (U c + Ws ) . Eq. 3-3
If the excavation is unsupported the relationship between the energy components can be
illustrated by an example with two identical specimens under constant load in a press
(Hedley, 1992), see Figure 3.2a. The stored strain energy (Um) in both samples would be
equal, see Figure 3.2b. If the load remained constant and then one of the samples was
removed instantaneously, the stress on the remaining specimen as well as the displacement
between the load platens would double. The stored strain energy in the removed sample (Um)
would be released, and the remaining specimen would have to carry the increase in stored
strain energy (Uc). The load platens would follow the displacement of the remaining
specimen, resulting in a change in potential energy (Wt). The various energy components are
shown in Figure 3.2b, with the following relationships:
U c = 2U m + U u Eq. 3-4
Wr = U m + U u = Wt / 2 Eq. 3-6
Wk = U u Eq. 3-7
20
where Uu is the increase in stored strain energy if the stress increase had occurred on an
unstressed sample (Hedley, 1992).
a) b)
Constant load
2σ1
Uu
One sample
removed σ1
Um 2Um
u 2u
Displacement
Figure 3.2. Energy components when one specimen is removed from a press with
constant load, after Hedley (1992).
When mining takes place in very small steps (incremental, infinitesimal), the limiting
conditions are:
Some interesting points that can be seen from these energy relationships are:
- If there is no support, all energy components can be expressed in terms of Um and Uu.
- If mining is done in small steps (infinitesimal), the process is stable and no seismic energy is
released. Some mines that implement incremental mining may still be experiencing
rockburst. This means that there are some other sources of energy being released, perhaps
due to inelastic conditions for instance fracturing of pillars or slip along a discontinuity.
- The change in potential energy is the driving force; if this can be reduced the other energy
components are reduced correspondingly.
- Support (backfill) is favorable in two ways; it reduces the change in potential energy by
reducing volumetric convergence of the stope, and it absorbs energy, which means that less
energy is available for release as seismic energy.
21
Seismic efficiency is a ratio used to describe rockburst potential. It is defined as the
proportion of energy released as seismic energy (Wk/Wr). The higher the ratio the higher the
rockburst potential.
To numerically study the energy components during incremental mining, Hedley (1992)
studied an unsupported stope in a vertical orebody. The orebody was 3 m wide and 30 m high,
mined in ten cuts. A pre-mining horizontal stress of 50 MPa was used. The change in the
seismic potential energy (Wt) for each cut is obtained by subtracting the total change in
potential energy of the previous cut from the present cut. Figure 3.3 shows how the energy
components change for each cut. The increase is linear as the mining progresses upwards. The
ratio Wk/Wr shows that 72 % of the total released energy is seismic energy.
Figure 3.3. Energy components per cut during mining of a vertical stope, from Hedley
(1992).
As an experiment, the topmost slice was divided into three 1 m slices to compare the seismic
efficiency between a 3 m slice and a 1 m slice. A 1 m slice should be a good approximation of
incremental mining. The reduction in slice height led to a decrease in seismic efficiency from
72 % to 59 % (Hedley, 1992), but is still far away from a zero seismic efficiency, i.e., that no
seismic energy is released. The equations used for calculation of the energy components can
be found in Hedley (1992).
22
3.2 Compressive stress and seismicity
Earlier it was assumed that stiff rock with higher uniaxial compressive strength and higher
Young’s modulus could store more strain energy than softer and more deformable rock. This
is not always true, which is illustrated in Figure 3.4. Brittle rocks usually have a steeper
unloading curve than soft rocks. In Figure 3.4 the area under the graph of the load-
displacement curves of both the brittle and the soft rock is approximately the same, which
means that both types of rock consume the same amount of stored strain energy during the
failure process.
Stiff side-
Soft side- wall rock
wall rock
σ1 σ1
σ1 σ1
σ1 σ1
Stiffness of
Stiffness of
surrounding rock, K
Stiff ore surrounding rock, K
Soft ore
Wk
Displacement Displacement
Figure 3.4. Stress-displacement relations for the combinations stiff rock – soft sidewall
rock and soft ore – stiff sidewall rock respectively.
23
When excavating in a stiff and brittle part of the rock mass where the surrounding rock has a
lower stiffness (stiffness of surrounding rock = K) than the unloading stiffness of the brittle
rock (the slope of the unloading curve), the failure process often becomes violent, see Figure
3.5. If the unloading stiffness is lower than the stiffness of the surrounding rock the failure
process will be non-violent, see Figure 3.5. This means that a strain burst can occur even if
the rock mass consists of only one rock type, since the unloading stiffness still can be higher
than the stiffness of the surrounding rock mass. In Figure 3.5 the response of the rock mass is
modeled by a rock specimen tested in a testing machine that is either stiff or soft. In the load
displacement curves, Um is the energy in the specimen, Us is the energy stored in the spring,
and Wk is the kinetic energy.
Figure 3.5. Model of the interaction between e.g., surrounding rock and ore (from
Hedley, 1992).
Seismic events are either compressive failures of the rock mass or slip on a discontinuity. One
pre-requisite for seismicity is a state of stress satisfying a yield criterion. Wawersik and
Fairhurst (1969) showed that stress alone is not sufficient to describe mechanical
consequences of a rock mass failure. Yield (failure) can be aseismic or seismic depending on
local strain energy density and the energy required to induce yield under the state of rock
mass confinement at the locus of rock failure. Wawersik and Fairhurst (1969) found after
24
testing samples in uniaxial and triaxial compression that the modes of response could be
divided into two classes. In the first class the rupture was stable, meaning that it took an
increase in load to decrease the load carrying capacity. In the second class the rupture was
unstable or self-sustaining, which means that the energy stored in the sample exceeded the
energy needed to cause fracture propagation. Another observation was that for a particular
rock material the classes of rupture could be separated as a function of the degree of
confinement during testing. At high confining stresses the unstable rupture could be
suppressed even for high axial loads, and a greater portion of the energy was dissipated as
post-peak yield.
N
K F
mg
τ
σn
A block of weight mg loaded with a normal force N is resting on a smooth horizontal surface.
The block is then loaded with a force F acting parallel to surface on which the block rests.
The force F is transferred to the block via an elastic spring with stiffness K. On the contact
surface between the block and the surface the normal stress σn and shear stress τ are acting.
As long as the block is immobile the shear strength will be µsσn, where µs is the static
coefficient of friction. The system is thus in equilibrium as long as
25
µ sσ n − τ > 0 . Eq. 3-9
τ = µ sσ n Eq. 3-10
we have an unstable equilibrium. The block will start moving for a small increase of the force
F, which corresponds to a small increase of the shear stress. A decrease of the normal stress
or the coefficient of friction will also cause the block to move. When the movement has
started the coefficient of friction decreases to µd and the block is accelerated by a force that
initially is equal to A(µs - µd)σn, where A is the contact surface between the block and the
horizontal surface. A stable equilibrium is reached when the shear stress equals µdσn.
The movement of the block (in Figure 3.6) as a function of time is shown in Figure 3.7b. Here
it is assumed that the free end of the spring has a constant velocity, s, in the direction away
from the block. If the end of the spring continues to move away from the block with a
constant velocity a new period of relative displacement between the block and the underlying
surface will occur at time t2.
Figure 3.7a) Stress-displacement curve for a block and b) movement of the block as a
function of time.
Figure 3.7a shows the stress-displacement curve for a block. Here the described course of
events is illustrated, where the relative movement starts when the shear stress equals µsσn and
stops when it has been reduced to µdσn. The curve that describes the movement of the block
can be non-linear even if the spring has a linear unloading curve K as illustrated in Figure
3.7a. The area between these two curves represents the energy that is transformed from strain
energy (potential energy) in the spring to kinetic energy for the block.
26
In the model described above, the shear stresses on the fault build up until they reach the
maximum shear strength of the fault, and then the earthquake occurs. This is termed stick-slip
behavior. If s, µs and µd are all constant the “earthquake” will occur at regular time intervals.
If the dynamic coefficient varies between the events, the time until the next event is
predictable but the size of the coming event is not. The time until the next event is
proportional to the amount of slip in the previous, which terms this model time-predictable
(Shearer, 1999). If the static coefficient varies instead, the time between the events cannot be
predicted. The amount of slip in the next event however, is proportional to the time since the
last event, and the model is hence called slip-predictable (Shearer, 1999). The problem is that
when both µs and µd vary neither the time nor the amount of slip can be predicted. An
assumption made in all these models is that the faults can be divided into segments that are
not interacting with each other. The Earth is however more complex than this since several
segments on a fault can interact and produce an earthquake, or that even two or more faults
can interact. The effects of these interactions have been modeled using the block-slider model,
Figure 3.8, where several blocks are connected by springs to each other and to a bar being
pulled at a constant speed, s. The slip of one block can set off adjacent blocks and lead to
large events. This type of model can produce a wide range of event sizes and a b-value (for
definition, see equation 5-12) close to that observed for real seismicity. The model can also
produce a chaotic order of events with no defined characteristic event or recurrence time.
The recurrence time of an earthquake can be modeled reasonably well as a Poisson process,
i.e., the probability of an earthquake at any given time is constant and independent of the time
of the last event (Shearer, 1999). Smaller earthquakes can be statistically modeled since they
occur frequently and have been cataloged for some time. Large earthquakes occur less
frequently so no statistical models can be tested due to lack of data. Even if long-term
prediction of earthquakes seems impossible, that does not mean that it is not possible to
27
predict events on a timescale from years to tens of years given sufficient knowledge of
stresses, strains and the local strength of faults. Prediction on the timescale minutes to months
is problematic. The focus so far has been on identifying definite precursors, i.e., anomalous
behavior observed prior to an earthquake. Possible precursors presented over the years are
changes in seismicity patterns, variation of the rate of emission of radon gas and
electromagnetic anomalies. However, the only definite precursor found so far is foreshocks,
which are events close in time and space to a subsequent main shock. The only problem is
that they do not always occur. One possible explanation to why earthquakes seem impossible
to predict was presented by Brune (1979). He proposed that large earthquakes start as smaller
earthquakes that start as even smaller earthquakes and so on. This makes it nearly impossible
to predict large earthquakes since even if every small tremor could be predicted how would
you decide which one would start a sequence of ever increasing events leading to a large
event? In his model he assumed that large parts of the fault existed at a state of stress below
that required to initiate slip, but that it can be triggered by nearby earthquakes or propagating
ruptures. He also suggested that precursory phenomena occur when the fault is close to the
yield stress. Support for his theory is that studies of the beginnings of earthquakes of different
sizes have shown no differences between large and small events. This means that from the
initial part of a seismogram, i.e., a graphical recording of an earthquake made by a
seismograph (Udías, 1999), it is impossible to say how large the event will become. This
seems to indicate that even prediction of small earthquakes is practically impossible.
28
4 ROCKBURST SOURCE MECHANISMS
In this chapter the different mechanisms causing rockbursts will be discussed. A summary of
the different terms used by other authors is also given. Many authors (eg., Cook, 1976;
Salamon, 1983; Ortlepp and Stacey, 1994) seem to agree on the basic definitions of the terms
seismic event and rockburst. A seismic event is the sudden release of potential or stored
energy in the rock. The released energy is then radiated as seismic waves. A rockburst is
defined as a mining-induced seismic event that causes damage to openings in the rock. There
are two general types of seismic events; those directly associated with stopes and those
associated with movement on major geologic discontinuities (Gibowicz and Kijko, 1994).
29
location of the failed pillar and the state of surrounding pillars and rock. The amount of
energy released by a pillar burst is much larger than that from a strainburst, so the radiated
seismic wave may cause damage in other areas such as shake-down of loose rock. The sudden
loss of support from one pillar causes stresses to be redistributed to nearby pillars, which in
turn may fail violently depending on how close they are to failure. A domino effect of pillar
failures may result, which may lead to collapse of that mining area.
Face burst is a form of strain burst and is caused by the accumulation of strain energy in the
fractured rock mass ahead of the face. Face bursts are accompanied by violent ejection of
material from the face into the excavated area.
30
4.2.2 Shear rupture
Shear rupture is a shear failure through intact rock, which occurs suddenly and causes
radiation of seismic waves and damage to nearby excavations. It requires a triaxial state of
stress and occurs when the compressive stresses ahead of a mining face exceeds the shear
strength of the rock. Another requirement is that the rock mass has to be free of joints. The
type of damage caused by shear rupture is the same as for a fault slip event.
4.3.1 Definition
The definition of a rockburst mechanism seems to differ among different authors. Some make
a distinction between the seismic source mechanism and the rockburst damage mechanism
(Ortlepp, 1997) while others use the term modes of failure (Kaiser et al., 1995), to describe
the same occurrences.
Ortlepp and Stacey (1994) suggested a classification scheme for seismic event sources in
tunnels, see Table 4-1. Their classification was based on first motions from seismic records,
the Richter magnitude of the events, and the postulated source mechanism.
Table 4-1 A suggestion for a classification scheme of seismic event sources in tunnels
(after Ortlepp and Stacey, 1994).
Ortlepp (1997) proposed a rockburst flow chart (Figure 4.1) to visualize the connection
between his definitions of source mechanism and damage mechanism, and to show a few of
31
the large number of factors contributing to the problem. First he describes event types (in the
horizontal direction) – strain burst, buckling, face crush/pillar burst, shear rupture and fault
slip. He then continues to define some important parameters describing the source
mechanism, such as magnitude, instability process, seismic signature and requisite conditions.
The instability processes are: spalling/buckling, euler-type instability, compression
slabbing/column crushing/dilation, and the stick slip behavior on faults. Seismic signature
describes the reaction to the event, e.g., if the event is implosive, if it is a shear event etc. The
last parameter is the requisite condition, which for a strain burst is that the failure surface
must be close to a free surface. If all conditions for the source mechanism are met a seismic
event takes place.
Ortlepp defined rockburst as a seismic event coupled with a damage mechanism and a
damage type. The damage mechanism depends on factors that determine the intensity of the
seismic impulse and factors that influence the site response. A few of the factors he mentions
as important for the intensity of the seismic impulse are amount of energy available, source
distance and dimension, and geological structures. Factors that influence the site response are
for example excavation geometry, characteristics of the surrounding rock (strength, brittleness
etc.), and characteristics of the existing support (strength, density, quality etc.). Finally he
specifies some damage types, for instance ejection, bulking, falls of ground, and shake-out.
32
EVENT strain burst buckling face crush, pillar shear rupture fault slip
TYPE burst
Requisite failure surface free surface stress > strength shear stress exceeds
close to a free >> lamina in destroyed
conditions surface shear strength shear strength
thickness volume
of rock at asperity
SEISMIC EVENT
intervening to cause
intensity of reflection, shielding,
seismic rate of liberation source ray path properties channeling or focusing of
impulse of energy dimension stress transient
ROCKBURST
33
4.3.2 Classification / categorization of rockbursts
This section will treat the different ways of classifying, categorizing and naming rockbursts
that can be found in the literature. In Table 4-2 some of the different methods are summarized
and in the following text each method is described in more detail.
Method of classification
Amount of rock Underlying
Distance Relation to
Author displaced, maximum mechanism
from /effect mining
magnitude, causing
inside mine situation and
amplitude on seismic
workings stress level
seismograph events
Scott, 1990 X
Scott et al., 1997 X
Johnston and Einstein, 1990 X
Knoll and Kuhnt, 1990 X X
Gaviglio et al. 1990 X
Krishnamurthy and Shringarputale, 1990 X
Bigarrre et al. 1993 X
Vervoort and Moyson, 1997 X
Wong 1992 X
Morrison and MacDonald, 1990 X
Hedley, 1992 X
Brummer and Rorke, 1990 X
Gill et al. 1993 X
Kleczek and Zorychta, 1993 X
Yi and Kaiser, 1993 X
Kaiser and Maloney, 1997 X
Arabasz et al. 1997 X
34
Table 4-3 Classification according to damage caused by seismic event.
According to these authors strain bursts are the cause of limited and localized damage, while
crush and slip bursts can cause extensive damage to drifts and stopes. Events exceeding 0.5
Richter magnitude, that displace more than 10 tons of rock into an opening, or have a peak
amplitude greater than 30 mm on a seismograph, are classified as large seismic events or
rockbursts. Microseismic or small seismic events are defined as events that displace less than
1 - 2 m3 of material into a mine opening, have a Richter magnitude less than 0.5, or result in
less than 30 mm displacement on a seismograph. These microseismic events generally occur
near drifts, haulage way or stopes, and cause popping, spitting (ejection of small rock
fragments) and spalling of the surrounding rock mass.
Distance from, or the location of the effect inside, the mine workings
Johnston and Einstein (1990), Knoll and Kuhnt (1990), Gaviglio et al. (1990), Krishnamurthy
and Shringarputale (1990), Bigarre et al. (1993), Vervoort and Moyson (1997) and Wong
(1992) all use the distance from the mining face or the location of the effect as a basis for their
classification of seismic events or rockbursts. The differences can be found in Table 4-4.
There is commonly one category “type 1” which is closely associated with active mining
faces. These events are a function of mining rate and are of low to medium magnitude. They
occur at or close to the mining face. The actual distance varies depending on author. The other
category “type 2” is only indirectly associated with mining activities, since it results from
regional stress redistribution caused by large working areas or whole mines. These are usually
located on a pre-existing zone of weakness or a geological discontinuity that can be located as
far away as 3 km. The coal industry has a slightly different approach – they treat pillar failure
as a separate group, called “Type 3” in Table 4-4. The pillars mentioned are stiff and become
overstressed because of the mining geometry.
35
Table 4-4 Classification of events according to distance from mining face.
Mining stage
Author
Development Extraction, active mining Mined out, whole mine
Morrison and Phase 1 events: in Phase 2 events: in or close Phase 3 events: some
MacDonald, 1990 development drifts to active orebody distance away
Inherent bursts: in
Hedley, 1992 Induced bursts: in remaining structures (pillars)
development openings
Morrison and MacDonald’s phase 1 events occur in development drifts when the pre-mining
stresses are high enough to cause failure by relaxation of stored strain energy. Phase 2 events
occur in or close to the orebody in the abutments (or pillars) due to redistribution and
concentration of stresses caused by the mining operation. Phase 3 events occur some distance
out in the walls of an extensively mined orebody, and are the result of regional stress
redistributions caused by the whole mine. The mechanism is always fault-slip, and it is not the
event itself that damages mine openings, but rather the vibrations caused by the event.
36
Hedley’s term for phase 2 and 3 events are induced bursts. Hedley also has a third type of
burst, fault slip, which can be either an inherent or induced burst, depending on the mining
stage at which it occurs.
These two mechanisms have commonly lead authors to define two classes of rockbursts. In
Table 4-6 rockbursts are grouped by using the type 1 and type 2 rockburst defined by Gill et
al. (1993). The type 1 bursts are those resulting from the dynamic loads imposed by fault-slip
events and the type 2 bursts result from failure of the rock mass itself. Sometimes there is a
third type defined, which is a combination of the two mechanisms, and is referred to as pillar
burst.
There are other authors using classifications of rockburst which do not fit into any of the
categories above. Petukhov (1990), for example, is the one who has found most classes of
rockbursts. He proposes five subgroups of rockbursts, that are differentiated by sound,
amount of dust dispersed, amount of shaking, location of damage in correlation to mining
stage and whether or not the mining operations are interrupted
37
38
5 SEISMIC MONITORING AND EARTHQUAKE SEISMOLOGY
5.1 Introduction
Seismic monitoring in mines is used to quantify the exposure to seismicity and is also a tool
to guide efforts to prevent or control seismicity. Mendecki et al. (1999) defined the following
objectives for monitoring of the rock mass seismic response to mining:
The information recorded by the monitoring system is hidden in seismic waves – to extract
the information for use in mine sequencing etc., methods used in earthquake seismology have
been adopted. Since seismic monitoring and the study of rockbursts and seismicity in mines
are applied forms of seismology, and many of the terms used in literature dealing with
rockbursts are borrowed from this field, this chapter will first describe seismic monitoring in
general and then some parameters from seismology used to describe seismicity are explained.
39
and the rate at which the rock is deformed during fracturing (Mendecki et al., 1999). To
describe a seismic event quantitatively, the time of occurrence and the location along with
either of the combinations seismic moment – radiated seismic energy, or seismic moment –
stress drop are needed. A seismic system can only measure the portions of strain and stress
that are associated with recorded seismic waves. When a number of seismic events within a
given volume and over a certain time have been recorded and processed, the changes in the
strain and stress regime in that volume can be quantified. This gives the opportunity to
validate results obtained from numerical modeling, where elastic parameters (E,ν) are
assumed to be constant within a given volume, making stress a function of strain (σ = Eε).
The strain and stress changes caused by seismicity, however, are independent (Mendecki et
al., 1999). Seismic strain in a given volume is proportional to the seismic moments (εs∝ΣM0),
and the stress is proportional to the ratio of seismic energies to seismic moments
(σs∝ΣEs/ΣM0).
The largest events that a mine seismic system has to record range between moment magnitude
Mw = 3 and Mw = 5, and the smallest are between Mw = -4 and Mw = -3 (Mountfort and
Mendecki, 1997). The range of frequencies that need to be recorded to make processing
meaningful is determined from the expected corner frequencies (see section 5.3.8) of the
events in the volume to be monitored. The part of the frequency spectrum where most of the
energy is released depends on the size of the event. For a large event low frequencies
dominate, i.e., frequency decreases with increasing size, and high frequencies are attenuated
faster with increasing distance away from the event (Jaeger and Cook, 1979). For the seismic
moment to be correctly determined, the requirement is frequencies five times lower than the
corner frequency of the largest event to be analyzed, and to correctly estimate seismic energy
the requirement is frequencies at least five times higher than the corner frequency of the
smallest event to be analyzed. Seismic events induced by mining have been shown to radiate
seismic energy from 10-5 J for microseismic events to 109 J for large rockbursts. The local
magnitudes corresponding to this energy release are ML = -6 and ML = 5, respectively (Jaeger
and Cook, 1979). The frequencies corresponding to the energy release ranges from less than 1
Hz to over 10 kHz. The choice of sensors to use in a seismic network depends on the desired
area of coverage and the sensitivity of the system. Two types of sensors cover the frequency
interval 1 Hz to 10 kHz, miniature geophones and piezoelectric accelerometers (Mendecki et
al., 1999). Geophones are suitable in sparse networks, where the distance between the sensors
40
is 1 km on average, for instance in monitoring several mining operations on a regional basis.
They are sensitive enough to record relatively distant events, i.e., they can record low
frequencies at large distances, and the risk of a large event occurring close enough to cause
clipping of the signal (i.e., exceeding of the maximum recordable ground motion) of more
than one sensor at a time is remote. Piezoelectric accelerometers are suitable in dense
networks, where the distance between the sensors is about 100 m on average. Many sensors
must be close together within the volume of interest, since the high frequencies they are
sensitive to quickly attenuate with distance. These accelerometers are good for mine-wide
monitoring.
Installation of both types of sensors should be in boreholes extending outside the fractured
rock surrounding an excavation. The sensor should be grouted in the hole to give good
coupling to the rock The grout should have similar acoustic impedance (product of density
and velocity of propagation) as the surrounding rock (Mendecki et al., 1999). The hole should
be completely filled around the sensor to avoid trapping acoustic energy. The sensors should
also be installed with a known orientation. For geophones it is important that the orientation is
within 5° of the vertical or horizontal otherwise they may not operate properly. Knowledge of
the orientation of the sensor also provides information useful for the localization of events,
and for the correct determination of the moment tensor.
41
can be seen as a step in between these extremes, where some events are on the same scale as
earthquakes and some are on the laboratory scale.
Strike or azimuth, α, is defined as the angle from north, while dip, β is defined as the angle
from the horizontal, see Figure 5.1. The slip vector is defined by the movement of the
hangingwall relative to the footwall and the rake, λ, is the angle between the slip vector and
the strike. The rake for vertical faults is defined with the hangingwall to the right of an
observer looking along the strike.
HW
Strike
N
α
Slip
Dip
FW λ
β rake
To describe the relative movement of the hangingwall with respect to the footwall, the terms
strike-slip and dip-slip have been introduced. Strike-slip means horizontal movement between
the fault surfaces while dip-slip means movement along the dip. If you are standing on the
footwall and the hangingwall is moving to the right or to the left, you have a right-lateral or
left-lateral strike-slip motion, respectively.
If the hangingwall is moving downwards in regard to the footwall, the fault is called a normal
fault. If the hangingwall is moving upwards, we have reverse faulting. Depending on the dip
angle the reverse fault can have different names. If the dip is less than 45°, the term is thrust
42
faulting, and if the fault is almost horizontal (dipping about 0 - 15) it is called overthrust
faulting, see Figure 5.2.
HW
HW HW
FW
FW FW
Figure 5.2 From left to right: normal faulting, reverse faulting and overthrust faulting.
Figure 5.3. Particle movement in P-wave (top) and S-wave (bottom), from Shearer
(1999).
The two types of waves propagate through the earth at speeds that increase with depth, see
Figure 5.4. The P-wave is always faster than the S-wave, and can propagate through all layers
43
of the Earth, but since the S-wave is a shear wave it cannot propagate through the fluid outer
core. Typical propagation velocities of P-and S-waves in granite near the surface of the Earth
are 5.5 km/s and 3 km/s respectively.
Figure 5.4. Variation of P- and S-wave velocities with depth beneath Earth’s surface,
from Kulhánek (1990).
Because of the velocity variations, the path of the seismic wave is a curve with the concave
side upwards (Arvidsson et al., 2000). This curved path is the shortest time-path through
Earth, see Figure 5.5. As a consequence of this, waves traveling to more distant measuring
stations penetrate to greater depths than waves going to nearer stations. The curved path is
also one reason that the P-waves of a near surface event (< 10 km depth) at teleseismic
distances usually have a downward first motion. The distance between the focus of the event
and the recording station is expressed as the angle between the source and the receiver, with
the point in the centre of the Earth. This means that 1° equals 111 km. Teleseismic distance is
a distance of more than 10° from the source.
44
Figure 5.5. Examples of propagation paths from a surface earthquake, from Kulhánek
(1990).
where G is the shear modulus at the source, D is the average shear displacement and A is the
area over which slip occurred. Sometimes the seismic moment can be calculated from
Equation 5-1 using field data, but most often it is estimated from seismic records. This,
however, requires more complicated calculations.
45
(Udías, 1995). This method represents the source by a system of body forces acting at a
point, and since the forces must represent the fracturing they are called equivalent forces. A
single force acting at a point can only result from external forces, otherwise momentum
would not be conserved. Internal forces representing the point source must act in different
directions. The simplest force system conserving momentum is the simple force couple or
dipole, Figure 5.6, meaning that there is no net moment.
If the forces are separated by a distance, d, in the direction perpendicular to the force
orientation, the momentum is no longer conserved unless there is another force couple
balancing the momentum. So a double-couple is a pair of force couples that ensures
conservation of momentum, that is, the net torque is zero, see Figure 5.7.
Figure 5.7. Unbalanced force couple (left) and balanced double-couple (right).
Force couples can be used to represent shear fractures, which is used in the formulation of the
moment tensor. To explain this we start with a simple point source. A point source can be
illustrated by an explosion on the surface of the sun. Since the sun’s surface is a liquid, the
explosion causes waves to propagate outwards in concentric circles (similar to those caused
by a stone hitting a water surface). The force diagram from the explosion can be seen in
Figure 5.8, along with an illustration of the first motion of the wave.
46
+
Figure 5.8. Force diagram from explosion and first motion of resulting wave (Andersen,
2001).
The motion of the wave can be described using force-couples. The three force couples needed
for the description of an explosive source are shown in Figure 5.9 along with the other six
force-couples forming the moment tensor.
3 3 3
2 2 2
1 1 1
M 11 0 0
0 0
3 3 3
3 3 3
2 2 2
1 1 1
Figure 5.9. The moment tensor with the three force couples describing an explosive
source circled (Andersen, 2001).
Any shear type movement can be described using combinations of the force couples in the
tensor. If we assume a vertical fault along which slip is initiated, P-waves will radiate
outwards from the point of initiation. The P-wave will form compressional (motion away
from the source) and dilatational (motion towards the source) quadrants around the source.
The compressional quadrants are denoted P+ in Figure 5.10 and the dilatational are denoted
P-.
47
+
P+ -
P-
Negative (down))
P-
P+
+
-
Positive (up)
Figure 5.10. Polarization of P-wave around the source of slip (Andersen, 2001).
The strongest movements are found in the middle of each quadrant, at a 45° angle to the fault
planes. If the two fault planes from Figure 5.10 are denoted X1 and X2 and the double-couple
force system causing slip on the faults are drawn in an X1 – X2 – coordinate system as in
Figure 5.11, then the double-couple can be equivalently represented by a pair of orthogonal
dipoles without shear. These dipoles are also called principal axes, and are denoted P- and T-
axes, where P is the compressional axis and T is the dilatational axis.
x2 x2
P-axis
x1 = x1
T-axis
In the double couple model there are two different fault planes corresponding to the same
seismic observations. The real fault plane is termed the primary fault plane and the other the
48
auxiliary fault plane. To determine which plane is the primary or auxiliary additional
information like aftershock locations or surveyed surface ruptures is necessary.
The moment tensor is a general representation of the forces that can act at a point in an elastic
medium. Although it is an idealization, it has proved to be a good approximation for modeling
distant seismic responses for sources that are small compared to the seismic wavelength
(Shearer, 1999). Since the moment tensor is symmetric it can be diagonalized, i.e., its
eigenvalues and eigenvectors can be calculated, and it can be further divided into an isotropic
part and a deviatoric part. The sum of the eigenvalues, m1 + m2 + m3, describes a volume
change at the source (Andersen, 2001), for instance
49
compressional quadrant is shaded, making the focal sphere look like a beach ball. Normal and
reverse faulting can be separated in these kinds of plots by noting if the center is black or
white. Normal faulting has a white center, and reverse faulting has a black center.
Figure 5.12. Focal spheres and their corresponding fault geometries, from Shearer
(1999).
50
time data is available the event can easily be mislocated. For example, if a fault separates two
blocks of crust with different velocities, the fault location tends to be mislocated off the fault,
into the block with faster velocity, see Figure 5.13. Also a poor distribution of stations around
the source can generate a large location error. If the earthquake occurs outside the seismic
network, the distance to the event is not well constrained, see Figure 5.14a, and the errors in
location form an ellipse with the long axis away from the seismic network. The location of the
event could be improved by having another station on the opposite side of the event.
Slow Fast
Figure 5.13. Location error caused by model errors, after Shearer (1999).
a) b)
Distant stations
Earthquake
Seismic network
Figure 5.14. Location errors caused by distribution of seismic array, after Shearer
(1999).
If only distant stations recording P-wave arrivals are available, the depth of the hypocenter is
difficult to specify since the takeoff angles of the rays are very similar, see Figure 5.14b. A
few stations at closer ranges help to pinpoint the depth of the earthquake, alternatively the
51
stations should record also the arrival of the pP-wave, i.e., the wave that from a deep focus
earthquake, propagated upwards and was reflected at the Earth’s surface, see Figure 5.15.
Figure 5.15. Examples of propagation paths from a deep focus earthquake, from
Kulhánek (1990).
Another way to enhance the location accuracy is to measure both P- and S-wave arrival times.
The S-wave travels at a lower speed, so the time difference between the arrivals can be used
to estimate the source-receiver range at each station. A rule of thumb is that the distance to an
event in kilometers is about 8 times the difference in S – P arrival times in seconds (Shearer,
1999).
7M 0
∆σ = , Eq. 5-2
16r03
52
which assumes total stress release. Equation 5-2 represents the uniform reduction in shear
stress producing seismic slip over an assumed circular fault, where M0 is the scalar seismic
moment and r0 is the radius of the circular fault.
Stress drops can vary considerably from event to event. For mine tremors the range is from
0.01 to 10 MPa (Gibowicz and Kijko, 1994). For earthquakes near plate boundaries (interplate
earthquakes) the stress drop is on average 3 MPa, while intraplate earthquakes (in the interior
of a plate) have an average stress drop of about 10 MPa (Shearer, 1999).
Richter introduced the local magnitude scale, ML, in the 1930’s. He noticed from plots of log
Λ (logarithm of amplitude) versus range that different earthquakes seemed to have a similar
decay rate. Richter defined the magnitude as the logarithm of the maximum amplitude Λ0 (in
micrometers) measured by a standard Wood-Anderson seismograph at a distance of 100 km
from the epicenter. The assumption behind the definition is that the ratio of maximum
amplitudes at two given distances is independent of the azimuth and is the same for all
earthquakes considered (Gibowicz and Kijko, 1994). This gives the relation
53
M L = log10 Λ(d ) − log10 Λ 0 (d ) Eq. 5-3
where Λ is the maximum amplitude of the current event at distance d, and Λ0 is the amplitude
of an M = 0 earthquake recorded at distance d (Udiás, 1999). Values of Λ0 for distances
between 10 and 600 km can be found in tables, e.g., in Udiás (1999). The advantage of the
Richter scale is that other subsequently derived magnitude scales have been related to it,
which makes comparison between events from all over the world easier. Disadvantages are
that it was defined for a southern California range - amplitude relationship and makes use of
an instrument rarely used today.
The so-called body wave magnitude, mb, is a more general magnitude scale for global
seismology, and is defined as
where Λ is the maximum amplitude, T is the corresponding period of the measured body
waves, and Q is an empirical function of range, h, and focal depth, ∆. The amplitude
measurements are usually performed on the first few cycles of the P-wave arrival. The body
wave magnitude scale has been calibrated to the local magnitude scale for small seismic
events in California (Shearer, 1999). Saturation starts at about mb = 5.5 and is total at about mb
= 6.5, which means that the magnitude rarely exceeds 6.5 even for very large events.
Another global magnitude scale is the surface wave magnitude, Ms, which can be defined as
where Λ is the maximum amplitude and T is the corresponding period of the Rayleigh wave,
ϕ is the epicentral distance in degrees, and 3.3 is a calibration constant. The period is usually
20 s. The surface wave magnitude can only be applied to near surface events, since the
amplitudes of surface waves are greatly reduced with depth. The saturation of the surface
wave magnitude begins at about Ms = 7.0. The body wave and surface wave magnitudes only
coincide at about Ms = 6.6, for smaller magnitudes mb is larger and for greater magnitudes Ms
is larger, see Figure 5.16. The relationship between the two magnitudes is
54
This relationship indicates that smaller earthquakes (Ms<6.5) are better measured by mb, since
the scale for Ms underestimates the size of small earthquakes (Udiás, 1999) and greater
earthquakes are better measured by Ms since that scale behaves well in the range 6.5 – 8.0, but
saturates strongly above Ms = 8.
5
M s, mb
3
Ms
2 mb
0
0 1 2 3 4 5 6 7 8 9
Ms
Figure 5.16. Relationship between surface wave magnitude, Ms, and body wave
magnitude, mb.
The difference between the scales and the saturation effect were the motivation for the
introduction of the moment magnitude, Mw, for great earthquakes by Kanamori. Later the
moment magnitude was used as a measure of earthquake strength. The formal definition of
Mw is made by Hanks and Kanamori (1979)
2
Mw = log M 0 − 10.7 Eq. 5-7
3
where M0 is the seismic moment measured in dyn cm (107 dyn cm = 1 Nm). The advantage of
this scale is that it is related to a physical property of the source through the seismic moment,
and that it does not saturate, even for very large events.
For mining-induced seismic events in Canada, the Nuttli magnitude scale is used. It is defined
as (Hedley, 1992):
Λ
M n = −0.1 + 1.66 log Ξ + log Eq. 5-8
ΓT
55
where Ξ is the epicentral distance in kilometers, Λ is half the maximum peak-to-peak trace
amplitude in the S-phase, Γ is the instrument magnification factor, and T is the period in
seconds. The relationship between the Nuttli and the Richter magnitudes varies, but in the
range of interest for rockbursts (between 1.5 < ML < 4.0) the Nuttli scale gives values that are
0.3 to 0.6 units higher than the Richter scale for the same event. Several authors have
formulated relationships for the two scales, and three of them are shown in Figure 5.17.
7
5
Richter magnitude, ML
1 2 3 4 5 6 7
Nuttli magnitude, Mn
Figure 5.17. Relationship between Nuttli and Richter magnitude scales, after Hedley
(1992).
which was defined by the Geophysics Division of Canada and is valid for 1.9 ≤ Mn ≤ 2.7,
56
defined by Boore and Atkinson in 1987 and valid for 4.5 ≤ Mn ≤ 7.0. All these relationships
are taken from Hedley (1992).
Large earthquakes occur more seldom than small earthquakes, and the same is true for
seismic events in mines. This magnitude-frequency relationship was described by Gutenberg
and Richter and can be written as
where N is the number of events with magnitude in the range M ± ∆M. Equation 5-12 can be
interpreted either as a cumulative relationship, if N is the number of events of magnitude
equal to or larger than M in a given time interval, or as a density law, if N is the number of
events in a small magnitude interval around M (Gibowicz and Kijko, 1994). The parameter a
is a measure of the level of seismicity. The parameter b describes the relative number of small
and large events in a certain time interval. The b-value typically varies between 0.8 and 1.2,
and b = 1 means that the number of earthquakes increases by a factor of 10 for every unit drop
in magnitude (Shearer, 1999).
57
where Et is the total seismic energy in ergs (1 erg = 10-7 J) and mb is the body wave
magnitude (Shearer, 1999). Gutenberg and Richter also developed the following relationship
between the local magnitude and seismic energy
where Wk is the energy in MJ (Hedley, 1992). A similar relationship has been developed for
the Ontario mines (Hedley, 1992) relating Nuttli magnitude and seismic energy
where ∆σ is the stress drop and M0 is the seismic moment. A velocity seismogram and the
corresponding frequency spectrum for the S-wave is shown in Figure 5.18. The lower
frequencies contain information on the coseismic strain changes, i.e., changes in strain caused
by seismicity, while the higher frequencies contain information on the coseismic stress
changes.
58
a)
b)
Figure 5.18. a) Velocity seismogram for a seismic event, and b) the corresponding
frequency spectrum of the S-wave (from McGarr, 1984).
Table 5-1. The abridged modified Mercalli intensity scale, after Bolt 1993.
59
Table 5-1. (continued).
60
Table 5-1. (concluded).
61
62
6 DESCRIPTION OF SEISMICITY
6.1 Introduction
This chapter summarizes some methods developed for the description, management and
control of rockbursts. A correct prediction would increase safety and decrease production
losses due to production stops and loss of drifts and stopes. The first three methods can be
used without having seismic records as a reference, and are commonly used to evaluate the
effects of a planned mining sequence. They can also be used for back-calculation of known
rockbursts if seismic records are kept. The other methods use numerical modeling in
combination with the seismic history of a mining step to try and predict areas where rockburst
may occur in the future. These methods are also used to identify hazardous areas in the mine.
All these methods strive at predicting rockbursts, – a fairly good prediction of the location can
usually be made based on seismic history in combination with stress analysis, but the time of
occurrence is not possible to predict.
∆A
63
A relationship between ERR, number of damaging bursts, and rock conditions for longwall
mines has been established by Jaeger and Cook (1979). A damage criterion has then been
formulated from this relationship (Jaeger and Cook, 1979; Hedley, 1992), relating the ERR
and the number of seismic events (Figure 6.2). The diagram shows that damage is negligible
when ∆Um/∆A is less than 15 MJ/m2 while extreme damage occurs if the release of energy
∆Um/∆A is larger than 100 MJ/m2.
Figure 6.2. Relationship between ERR and number of seismic events (modified after
Jaeger and Cook, 1979)
The ERR gives a hint of the rockburst potential within the mine workings. The ERR is closely
related to the volumetric closure, which both can be easily calculated for elastic conditions
but they give little information of the behavior of fractured rock masses (Spottiswoode, 1990).
ERR was (1990) used widely for designing mining layouts in deep mines with the purposes of
reducing seismicity, and is still used today in some mines.
64
6.2.2 Excess Shear Stress (ESS)
The quantity Energy Release Rate is only a measure of the rockburst potential of a mining
stope and does not give any information about the risk of seismic activity along large
geological structures. Slip on such structures has from South African experience been
observed to cause substantial damage in large parts of a mine. In the Sudbury mining area in
Canada it is assumed that all seismic events causing Richter magnitudes greater than 3.0 are
so called mining induced earthquakes.
Ryder (1988) developed a criterion based on the excess of shear stresses along a discontinuity
(“Excess shear stress”). The model used is the stick-slip model described in Chapter 3.3. At
the point where slip occurs, the balance between shear forces and counter forces can be
expressed as
τ s = c s + µs σ n Eq. 6-1
where τs is the static shear strength, cs is cohesion, µs is the static coefficient of friction and σn
is the normal stress.
When slip has been initiated, the dynamic coefficient of friction and the normal stress
determine the shear resistance. The cohesion is assumed to become zero when slip has been
initiated. The difference between the static and dynamic shear stress equals the excess in shear
stress (ESS), τe
Ryder (1988) suggested a preliminary value of the critical excess shear stress in the order of 5
to 10 MPa for existing weakness planes while the corresponding value for failure through
intact rock would be 20 MPa. A typical value of the static coefficient of friction is 0.6, which
corresponds to a friction angle of 30o. The values stated here are valid for South African rock
conditions.
Ryder’s concept for slip on a fault is illustrated in Figure 6.3. The fault plane is assumed to
have an uneven static shear strength distribution caused by irregular contact surfaces and
varying frictional properties, which is represented by the uppermost curve in Figure 6.3.
When slip is initiated the shear strength instantaneously drops to a smoother shear strength
65
distribution with lower magnitudes. If the static shear stress distribution at any point becomes
high enough to equal the static shear strength, slip is initiated at this point causing an
immediate decrease of the shear strength to the level of the dynamic shear strength (stress
drop). High stress concentrations on the boundary of the slip surface cause the slip to expand
into the area around the point at which slip was initiated. The area of slip propagates along the
fault to areas of lower excess shear stresses. Since the process is dynamic, slip can also occur
in neighboring areas where the static shear strength is lower than the dynamic.
In South Africa ESS is used to predict the risk of slip on faults as follows: Using numerical
modeling of different mining sequences, normal and shear stresses along a given fault are
calculated. If the static shear stress exceeds the static and dynamic shear strength, the area
where slip may occur is calculated, see Figure 6.4. As a rule of thumb it is assumed that if the
maximum excess shear stress is of the order of 5 to 10 MPa slip occurs and the displacement
along the fault is calculated. To obtain better values for calculating the critical excess shear
stress it is assumed that a fault without filling has a friction angle in the interval 28 - 30o
(which is close to the base friction angle) while the friction angle of a clay-filled fault is in the
interval 10 - 15o. In areas with high seismic activity some kind of seismic monitoring is
performed. With these instruments the stress drop due to slip on a fault can be determined.
Based on these measurements the dynamic coefficient of friction for a specific fault can be
66
determined, thus gaining better data for predicting the risk of a seismic event in the form of
fault slip.
Figure 6.4. ESS contours and faults where there is a risk of slip.
Figure 6.4 shows an example of how the technique described above has been used to predict
the risk of slip on faults dipping 70o. The tangents indicate the most probable point of slip
initiation for three possible maximum stress drops, τe = 10, 5 and 2 MPa. The length of the
line indicates the associated length of rupture, and next to each line are the estimated
magnitudes. Ryder showed that for the combination of primary stresses and mining method
(longwall mining) representative of South Africa, faults dipping 70o or 130o are easiest to
activate. To estimate the magnitude of the potential seismic event, in this case slip along a
part of a fault, Equation 5-1 is used.
67
M 0 = G ∫ ∆ε st dV Eq. 6-3
where G is the modulus of rigidity or shear, and ∆εst is the static shear strain change
associated with the seismic event. Using τ∆εst = G∆τ Equation 6-3 can be rewritten as
where the right hand side of the equation is termed Volume Excess Shear Stress (VESS).
Some assumptions are connected to this definition. First, it is assumed that all positive ESS
values are reduced to zero through shear slip during seismic events. This means that after the
slip has taken place there are no excess shear stresses left on the fault. Contributions caused
by integration outside the volume of positive ESS are assumed to be unimportant and are
neglected. VESS is calculated by integration of ESS over the positive region. VESS was tested
by Spottiswoode (1990) in two ways. First, the maximum seismic moment release for a given
region at any time M0,max was calculated with Equation 6-4, resulting in an estimation of the
magnitude of the largest event possible in that region, assuming that the largest event causes
half of the volumetric change. Secondly, the cumulative seismic moment during a given time
period was determined from changes in VESS. It was found that the cumulative seismic
moment calculated using VESS did not correlate well with the observed value, as it was about
five times smaller than the observed value. The VESS method, however, quite accurately
estimated the magnitude of the largest event.
( Pi / Pj ) msd
I D ( Pi / Pj ) = Eq. 6-5
( Pi / Pj ) aver (T =t )
68
where ID(Pi/Pj) is the departure index of a given ratio of the parameters Pi and Pj, which are
the seismic parameters used for the analysis. (Pi/Pj)msd is the ratio of seismic parameters for a
given seismic event, and (Pi/Pj)aver(T=t) is the ratio of seismic parameters averaged over a
number of events during a selected time period. Examples of parameters that can be used in
the analysis are seismic energy of P-and S-wave, apparent volume, seismic moment, seismic
radius, b-value, corner frequency of P- and S-wave, and static stress drop. The seismic
monitoring system records values of these parameters, and an algorithm calculates the
departure index. The advantages of the method are that
The departure index can also be used to evaluate the rockburst hazard. The rockburst hazard
indication (RHI) is expressed as
where
1 if I D ≥ I D ,crit
ΨD ( Pi / Pj ) = . Eq. 6-7
0 if I D < I D ,crit
This means that when ΨD = 1, there is seismic risk and that departure index contributes to the
rockburst hazard indication, and when ΨD = 0 there is no seismic risk, and consequently this
departure index does not contribute. The value of the rockburst hazard indication is after
calculation compared to a critical value, which is based on experience from the mine in
question. When the rockburst hazard indication exceeds the critical value, it means that a
rockburst is imminent (Poplawski, 1997).
69
6.3.2 Cell evaluation method - Modeled Ground Work (MGW)
An increase in the number of incidents (damage to mine openings and human injuries) related
to seismicity in Australian mines has lead to a need of a method for managing seismic hazard.
Such a method was proposed by Beck (2000) and it compared different excavation options in
terms of a quantitative assessment of seismic hazard. This assessment was made using a
three-dimensional elastic boundary element model of the host rock mass. Parameters from the
analysis could then be compared to records of observed seismicity to provide probabilistic
relations between seismic event occurrence and event strength. Beck suggested that the
parameters that should be used for event occurrence relations and event strength estimates
were Factor of Safety against seismic failure for seismic event types determined by back
calculation and Modeled Ground Work (MGW) or Local Energy Release Density (LERD).
Both MGW and LERD are numerical models for comparing the load-deformation state of a
volume of rock before and after an event. This is done by simulating the event as a change of
material properties within the volume. These methods are described below.
Beck and Brady (2002) stated that a seismic event can occur if the conditions of a stress state
sufficient for failure and an energy condition such that failure is violent, are met. They
presented the hypothesis that the potential for rock mass failure and strain energy density
distribution are the most relevant physical parameters affecting the occurrence of induced
seismicity. The cell evaluation method was developed for evaluating the magnitudes of the
controlling parameters throughout the mine domain. The method requires that the whole
volume of influence of an historical mining step can be discretized into regular test volumes
(finite cells). Another assumption is that the evolution of the mining geometry and field
observations of seismic events (location and magnitude) are available in the volume of
interest. The controlling parameters (yield potential and strain energy state) are calculated in
each cell. By comparing and correlating the parameter values in the cells with the rate of
occurrence of seismic events (in reality) a multi-variate probabilistic relation between the
parameters and the seismic outcomes of a step in an excavation sequence may be defined.
70
of seismic events resulting from future mining activity. The probability of an event X of
magnitude x is:
Beck and Brady propose in their model the use of a “factor of safety” to evaluate the stress
parameter (Q1) for different types of events and to provide a quantitative measure for planned
mining. The factor of safety is defined as the ratio of rock strength to rock stress. For slip on a
discontinuity with zero cohesion and using Coulomb’s failure criterion, the factor of safety is
defined as
σ n tan φ
ST = Eq. 6-9
τ
where σn is the normal stress acting on the discontinuity, φ is the friction angle and τ is the
maximum shear stress.
Beck and Brady (2002) developed the definition/specification of the strain energy parameter
(Q2) after studying the results by Wawersik and Fairhurst (1969) regarding the energy of
specimen rupture. Beck and Brady specified that Q2 ideally should uniquely describe the state
of the factors that control rupture violence. Possible parameters are ERR, ESS, LERD and
MGW. To describe their model Beck and Brady used MGW, formulating the probability of an
event X of magnitude x as
The functional relation f is found by establishing how often different combinations of ST and
MGW result in seismic yield, and thus the probability that these combinations result in future
seismicity can be calculated.
71
Beck and Brady applied the method to mining steps in two different mines and found that it
was possible to establish relations between seismic event strength, event probability, MGW
and ST. They went on to show how these relations could be used to quantify the seismic risk
of a planned excavation step. This is done in four steps:
1. A regular volumetric grid is established around the likely volume of influence of the
proposed mining geometry.
2. At the test cell centers Q1 = ST is calculated, the potential event types in each test cell
is recorded, and Q2 = MGW related to the specific type of seismic event associated to
the value of ST is determined.
3. The expected probability of event occurrence within the cells is determined using plots
derived empirically for the particular mine site or stope block.
4. When seismically susceptible cells are identified, MGW is compared to the derived
seismological relations to estimate the event strength.
The procedure was evaluated for cases from the same mines as used previously and the
method gave good agreement between assessment of event probability and later observations
of event occurrence. Event strength is another important issue for a mine and here also there
was good agreement between predicted results and field observations. The conclusion from
the case studies is that the method produces usable relations for management of seismic risk,
and suggests that the cell evaluation method has a potential use as a framework for evaluating
seismic hazard generally.
1 n n n
U1 = ∑
2 i =1
t xi u xi + ∑ t yi u yi + ∑ t zi u zi
,
Eq. 6-11
i =1 i =1
72
where n is the number of elements on the surface of the test volume. The strain energy after
the episode of rock mass degradation (post-event) is:
1 n ' ' n n
' '
U2 = ∑ xi xi ∑ yi yi ∑ zi u zi
2 i =1
t u + t ' '
u + t Eq. 6-12
i =1 i =1
MGW is defined as the area below the model characteristic curve, see Figure 6.5, and is
expressed as:
3 n
MGW = ∑ ∑ 12 (t ij + t ij' )(u ij' − u ij ) . Eq. 6-13
i =1
j =1
Model
characteristic
73
material with reduced cohesion and friction is a more realistic substitute. When the
calculations of stresses and displacements are completed, the original material is replaced, and
then the process is repeated for all other test blocks at each mining step. The total energy
released by the system is found by first integrating the stresses over the area (to obtain the
load) and then to integrate the load over the displacements resulting from the substitution of
the alternate material. Finally the total energy released is divided by the test block volume, so
the LERD will be expressed in MJ/m3. This way of determining LERD means that it is a local
rock mass characteristic that will vary in magnitude from one place to another, and also that
the value at any location will change as mining progresses. This means that for instance a
pillar can pass in and out of the burst prone state as mining proceeds.
Stage I Stage II δ
The LERD-method was tested on some cases taken from INCO’s Creighton Mine (Sudbury,
Canada), where stress state and LERD were calculated as mining progressed. The back-
calculation was used to verify the method. The results indicated that locations that had a
critical level of LERD and stress, were locations that were reported to have burst. By studying
the distribution of LERD, the most likely location of a rockburst could be identified, and also
the type of the anticipated failure was indicated as progressive or violent (Wiles, 1998).
74
mines that responded were open stoping (52 mines), entry methods, i.e., cut-and-fill and
room-and-pillar, (12 mines), sub-level caving (7 mines) and block caving (2 mines). The
survey had 148 questions, which included questions regarding rock properties, geology,
stress, orebody, mining method, mining geometry, seismicity, methods of managing
seismicity, and related rock mass damage. The data collected was used to train an artificial
neural network to replicate the results of the survey. Seismic hazard is defined as the
likelihood of occurrence of events of a certain magnitude (Hudyma, 2004). The seismic
hazard scale (SHS) uses three common mine seismicity parameters to rate the relative
occurrence of seismic events of various sizes in a mine. The parameters are
The SHS assumes a b-value = 1, which means a 10 fold increase in number of events for each
unit decrease in magnitude. The SHS has been used to investigate seismic hazard for events
up to ML = 2 - 3. Events of larger magnitudes (large fault-slip events) may not follow the
scale, since their b-value may be less than one. Also the failure mechanism of large events
may be different than for smaller events, so using the SHS methodology could lead to
misleading estimates of seismic hazard (Hudyma, 2004). Depending on the mining method
the applicability of the SHS varies, due to the distribution of mining methods in the
participating mines. This concerns mainly sub-level caving mines, where it should be
remembered that only seven mines responded to the survey. The SHS is not valid for block
caving mines since only two mines completed the survey. Another aspect is that the SHS is a
measure of past seismicity in the mine, and if the rock mass conditions change or if the
mechanisms or driving forces of future seismicity change, forecasting using the SHS may be
misleading.
Hudyma (2004) developed a method to calculate the seismic hazard for different mines and
orebodies (some mines provided information on more than one orebody), and then the SHS
was plotted against several geotechnical, mining and geological parameters. The resulting
correlations between SHS and the different parameters are summarized in Table 6-1.
75
Table 6-1. Summary of correlation between SHS and geotechnical, mining and
geological parameters (Hudyma, 2004).
The advantage of the SHS is that a mine can evaluate the seismic hazard without having data
from a seismic monitoring system. The SHS also gives a reasonable estimate of the maximum
size of events that can occur in the mine (Hudyma, 2004).
76
7 PREVENTION AND CONTROL OF SEISMICITY
The cause of rockbursts is (in many cases) a combination of stiff rock and stresses high
enough to exceed the strength of the rock. It should be noted that strain energy is accumulated
in a volume of rock, but that the energy is dissipated along surfaces within the volume (Blake
et al., 1998). The potential of violent failure is also higher in homogenous rock, i.e., rock with
less natural discontinuities or with little variation in mineralogy. An inhomogeneous rock
mass is more likely to develop micro-fractures, or shear along pre-existing joints, leading to
lower stiffness and greater energy dissipation (Blake et al., 1998). To alleviate violent failure
one can either make the rock less stiff by creating shear along existing weakness planes or
decrease the stresses acting on the piece of ground subject to bursting.
To soften the rock or transfer the stresses several approaches can be used. Changing the
mining layout can ensure that pillars are not overstressed. Changing the shape of an opening
can also decrease stress concentrations in unfavorable locations on the boundary. If it is not
possible to change the shape of the openings or the mining layout there are other methods of
accomplishing a reduction in the potential for violent failures, which may be less efficient but
yet sufficient.
It was noted above that to reduce the potential for violent failure, either the stiffness of the
rock should be decreased or shearing on existing fracture surfaces should be promoted. There
are two ways of accomplishing this by destress blasting. The first alternative is to blast as
heavily as possible without damaging the excavation too much. The idea is to soften a certain
region by creating as dense a zone of microcracks as possible. The softening will lead to an
alteration of mechanical response of the rock mass from elastic-brittle to plastic deformation
(Blake et al., 1998).
77
Toper et al. (1997) uses the second approach and states that preconditioning does not cause
formation of any new fractures ahead of the face but instead leads to slip on pre-existing
fractures due to the high gas pressure from the explosion. The fractured rock mass that results
from the preconditioning blast forms a “protective cushion” ahead of the face. If a seismic
event would occur at some distance from the preconditioned face the dynamic tensile wave
from that event would not cause as much damage as for a normal face. The effect of a
preconditioning blast is localized in space, so to avoid face bursting on a specific panel, that
particular face must be preconditioned. The effect is also limited in time. Since the
mechanism of preconditioning is one of stress transfer as a result of induced deformations in
the fractured rock ahead of the face, the fractured zone is still capable of carrying high loads.
It is therefore possible that future mining can transfer stresses back toward the previously
preconditioned face and that a face burst may occur. An example of this from Western Deep
Levels South Mine (South Africa) is described by Toper et al. (1997). Mining on a panel (E)
that had been pre-conditioned was stopped, while mining continued on neighboring panels (C
and F). Two weeks later a magnitude 1.1 event occurred in the vicinity of panel E, triggering
a face burst on that panel, which destroyed a supporting pack.
As a general rule destressing of development drifts is often successful, and many mining
companies have developed their own standards, which cover drill patterns, explosives,
charging and blasting (Blake et al., 1998). Destressing of pillars is more difficult, since the
effect of the blast is not well understood. Destressing of one pillar also transfers stresses to
neighboring pillars, which in turn may become to highly stressed.
78
Face-parallel method
A good mining layout is shown in Figure 7.1, where drilling and blasting of the
preconditioning holes has been integrated into the production cycle. Each panel has a
maximum length of 20 m, and the panel lead has been minimized. Panel 2 has been mined to
the limit (i.e., the previous preconditioning hole) and the next hole is drilled at about 5.5 m
from the current face.
Unmined Length = 20 m
Panel 1
5.5 m
lead
Previous pre-
conditioning hole
Panel 2
New pre-
conditioning hole
Panel 3
Figure 7.1. Preconditioning layout in an overhand mining sequence (after Toper et al.,
1998).
The preconditioning hole should be drilled parallel to the reef plane along the reef/footwall
contact. The hole is drilled in the fractured zone ahead of the face and should extend into the
next panel. To ensure proper destressing of the entire face it is important that the holes are
parallel to the face. It is also important that the hole is drilled in the correct position ahead of
the face. If it is drilled too far away drilling will be difficult because of high stresses and
blasting could lead to stresses being transferred back towards the face. The actual position of
the hole is site-specific and depends on both stress state and the size of the charge to be used.
79
offset horizontally by 0.5 m from the previous ones to avoid drilling into old sockets. Each
production blast moves the face forward about one meter so every three production blasts the
preconditioning holes are drilled in the same position. Research referenced by Toper (1997)
has shown that the spacing between the holes should not exceed 3 m. This has been concluded
since the radius around the holes affected by a trial blast was about 1.5 m. Trial and error has
also lead to the conclusion that a length of 3 m and a diameter between 36 mm and 40 mm is
the optimum.
0.5 m
Day 3
3m
Day 2
Day 1
3m
Direction of
face advance
Figure 7.2. Face perpendicular preconditioning holes, after Toper et al., 1997.
The preconditioning holes in Figure 7.2 are drilled at mid-height of the stope to reduce the
risk that any of the production holes intersect the preconditioning holes, see Figure 7.3. The
preconditioning holes are charged for the bottom 2 m and top-primed and then the remaining
1 m is stemmed using for example bentonite or angular sand.
Hangingwall
Preconditioning hole
Production
holes
Footwall
1m 2m
Figure 7.3. Cross sectional view of face perpendicular layout after Toper et al., 1997.
80
Both the face parallel and the face perpendicular methods have lead to a decrease in the
number of face bursts and have improved working conditions. Both methods also improve the
drilling rates for the production holes, since it is faster to drill through fractured rock than
intact rock.
7.1.2 Sweden
Destressing has been used in several Swedish mines, among others Näsliden, Laisvall,
Malmberget and Kristineberg mine. The two first mines have now been closed, but it is still
interesting to see how it was done there. Destressing performed by Boliden up until 1988 was
summarized by Krauland and Söder (1988).
Näsliden mine
The first attempt was made in Näsliden mine in 1978. The mining method used was cut-and-
fill and the stress situation was characterized by high horizontal stresses causing extensive
spalling of the roof of the stopes, often in combination with minor rockbursts. The orebody
was tabular and steeply dipping. The destressing attempt had two objectives, namely
- to show and quantify the destressing of a cut-and-fill stope by a correctly placed destressing
slot, and
- to find a good explosive and to examine the crushing of the rock around the blasthole.
To determine the correct placement of the destressing slot, numerical models were analyzed
using a finite element-program. The purpose of the destress blast was to decrease the stiffness
of a part of the rock mass close to the excavation, so that the stresses were redistributed away
from the boundary of the excavation. The model showed that if the ratio of the stiffness of the
destress slot to the stiffness of the in-situ rock was 2.5% - 10% the horizontal stresses in the
roof decreased by 37 - 60%, see Figure 7.4.
81
Before Destress slot
destressing
Reduction of Young’s modulus of rock
in slot
E slot
E original
without destressing
with destressing
destress slot
Figure 7.4. Modeling of the effect of destressing of the roof in a cut-and-fill stope, by a
destressing slot in the footwall, after Krauland and Söder (1988).
Field trials were then performed in the #3 stope, to validate the model results. The slot was
blasted in the footwall rock to destress the roof. The destressing of #3 stope failed because the
rock was not sufficiently crushed around the blastholes. The reason for this was that the
explosive had too low detonation velocity, and that the slot had been placed in a ductile
chlorite. After some investigation it was determined that an explosive with a high detonation
pressure should be used. Some conclusions could be drawn from the failed attempt for
destressing to be successful:
82
- the stability problems should be bad enough for shotcrete and rock bolting not to be
sufficient,
- the destress slot should be placed far enough into the rock mass so that the blasting does not
damage the roof or walls of the excavation, and
- several destressing holes should be placed close to each other so that a slot of cracked rock
is formed.
In 1988, destress blasting was attempted in #5 stope because of high horizontal stresses
leading to rockburst problems and intensive spalling of the roof. The high stresses were
caused by a diminishing sill pillar. To get a more stable roof destress blasting was tried during
mining of the last but one slice. The height of the sill pillar at that time was 10 m. The
destress holes were blasted together with the production rounds. During the first four rounds
there were no noticeable effects, but after the fourth round the rockbursts stopped and the
extent of spalling of the roof decreased.
Laisvall mine
Laisvall mine was a room-and-pillar mine characterized by high horizontal stresses despite a
relatively low depth, about 220 m (Engberg, 1989). The vertical stress is gravitational and
approximately 6 MPa at this depth, but horizontal stresses of between 20 and 25 MPa have
been measured, which leads to a horizontal to vertical stress ratio of 3.3 – 4.2. The orebody is
tabular and subhorizontal. The mining layout of the trial area, Nadok 1400, is shown in Figure
7.5.
1400-drift
1500-drift
Figure 7.5. Wedge-shaped mining area in Nadok 1400, after Engberg (1989).
83
The 1400-drift was excavated first, followed by the 1500-drift which was about 15 m behind.
Mining perpendicular to the two main drifts was carried out so that the mining front became
wedge-shaped. Spalling failures in the roof took the form of an inverted keel, see Figure 7.6,
and occurred mostly during excavation of the first drift. The depth of the failures was 0.8 –
1.5 m. Thus, it seemed that the second drift was excavated in a destressed area. Bolting was
not sufficient to stop the spalling and secure the stability of the drift; hence it was decided to
try destress blasting.
The destress blasting was conducted using three parallel holes that were drilled together with
the production round. The holes were located in the right abutment with a c/c of 1.4 m. The
average direction is 42° away from the drift and about 25° upwards, see Figure 7.7.
Destress hole
42°
25° Destress
holes
1.4 m
3m
The bottom part of the holes (0.6 - 0.7 m) were charged with 1 kg of explosives, and the holes
were blasted with the production round. Drilling into the area that was blasted gave an
approximate size of the cracked zone of 2-2.5 m width and 1.5 m height. The cracked zone
was most likely affected by the anisotropy of the clay schist in the roof so that the zone of
damage was not perpendicular to the axis of the hole. Rather it had propagated along the
weakness planes of the clay schist, see Figure 7.8.
84
Figure 7.8. Rotation of cracked zone, after Engberg, 1989.
Damage mapping of the roof was performed in two areas. Destress blasting had been
conducted in one area, and the damage here was less extensive regarding both area and depth
of failure. In a few places the destress blasting had not been performed as planned and in
these areas the roof damage had increased (Engberg, 1989). This indicated that the destress
blasting was effective in reducing the stresses in the roof.
Malmberget mine
During excavation of the main level at 815 m (1985) the drifting was slow due to severe
seismicity problems. The layout of the main level is shown in Figure 7.9. The main rocktypes
in the area are red- / redgrey leptites, which are sometimes gneissic in structure and can also
sometimes be weathered. Biotite and granite occur as veins. The transportation drift had an
area of 60 m2, and the seismicity led to spalling of the roof up to a height of about 4 m above
planned roof level. The spalling resulted in either a sloping roof or a “church” forming, see
principal sketch in Figure 7.10. To increase the safety and also increase the drifting rate, trials
with destress blasting were conducted.
85
Figure 7.9. Layout of main transportation level 815 m (Borg, 1988).
The first step was to observe the seismic events and try to correlate them with geology. It was
found that the worst problems occurred in aplite and red leptite, which are the two rock types
that have the highest uniaxial compressive strengths, 194 MPa and 178 MPa, respectively.
These rocktypes resulted in the largest overbreak (Borg, 1988). Water bearing zones also
caused a lot of seismicity, but no problems with overbreak. In step 2, a numerical analysis of
the stress situation around the drift, both with and without a destressing slot, was performed.
a) b)
9.8 m
Figure 7.10. Principal sketch of a) sloping roof profile and b) churching (Borg (1999).
86
The slot had a height of 1 m or 2.7 m and a length of 4 m, see Figure 7.11, and was modeled
as a zone with lower Young’s modulus (from 12 to 28 GPa) than the surrounding rock (40
GPa). The numerical model assumed that the rock was isotropic and elastic and that there
were no dynamic sequences.
9.8 m
Figure 7.11. Principal sketch of drift with a 2.7 m high slot used in numerical modeling.
The numerical modeling showed that the slot decreased the stresses in the center of the roof,
and a high, long and soft slot was most advantageous. Taking both these results and previous
destressing experiences (from e.g., Boliden mines) into account it was decided to try destress
blasting in one of the sidewalls with the layout in Figure 7.12 (Borg, 1988). The idea was to
redistribute the stresses near the boundary of the excavation farther into the rock, and to avoid
damaging the rock close to the boundary.
Slot
Drill holes
9.8 m
After a few rounds with this layout it was decided to drill only three holes to avoid damaging
the boundary rock too much. Hence destressing of both sides of the drift was done using the
same layout on both sides. The original drill plan is shown in Figure 7.13, and Hole 0 is the
one that was subsequently removed. The distance between the holes was 1 m, measured
87
perpendicularly to the hole axis. The holes were charged with one kilo of explosives, packed
into the hole bottom.
Hole 3
Hole 2
Hole 1
Hole 0
The result of the trial with destress blasting was that the seismicity and overbreak decreased
noticeably in the sections that were destress blasted.
88
radial flow), hence the area that had enough pressure to initiate slip was rather small. A slip
radius of about 2 m is consistent with the small event magnitudes (Board et al., 1992).
7.3 Reinforcement
7.3.1 Introduction
Reinforcement is used in almost all underground mines to stabilize mine openings and secure
the safety of the personnel. The reinforcement standards differ between countries, companies
and mines, depending on different cultures and requirements. Most mines probably start out
reinforcing to keep wedges in place, but as the mine grows in size, age, and depth, other
problems often come up. One of the problems that seems to occur with increasing depth is
seismicity and rockburst. Since seismicity is a dynamic problem the reinforcement used for
static stabilization may not be sufficient. If the problem is localized to only a portion of the
mine, destressing may be attempted to relieve the stresses. However, if destressing is not
sufficient to entirely prevent rockbursts from occurring, reinforcement has to be used to
stabilize the rock around the excavation.
89
may be preferable. The holding elements may in some cases be required to absorb energy to
decelerate ejected blocks, but in other cases they may only be required to move without
absorbing energy (Kaiser et al., 1995). A support system is made up of separate elements that
work together to perform the reinforcing, retaining and holding functions described above, see
Figure 7.14. This requires a good design of the connections between the elements, for
example shotcrete and bolts.
Figure 7.14. The primary functions of support elements, from Kaiser et al. (1995).
90
Table 7-1. Characteristics of different supports and what functions they fit best for, after
Kaiser et al. (1995).
Kaiser et al. (1995) estimated that a holding element that could accommodate deformations of
up to 300 mm at loads greater than 100 kN would be best for burst-prone ground. They tested
several holding elements and found that none quite matched these requirements. The bolts
that best matched the requirement included cone bolts, yielding Super-Swellex bolts and cable
bolts that allow slip. Retaining elements are in general weak and soft, with the exception of
shotcrete. At small wall deformations (> 50 mm) corresponding to the peak load capacity of
standard non-yielding bolts, mesh provides little support. When the rock supports itself or
when competent rock arches form between the rock bolts the weight of rock that the mesh has
to support can be quite small. However, if key blocks in a stable arch start to unravel, the arch
looses its stability and the support has to carry the weight of the rock. To help the rock
support itself it is important to restrict the movement of key blocks. In this regard a mesh with
stiff load-displacement is a better support since it will minimize the deterioration of the rock
above. For this reason it is better to use a diamond bolt pattern than a square because the
distance between the bolts will be smaller and that also helps to minimize loosening of the
rock mass. If mesh is combined with shotcrete the initial stiffness increases, and even at large
displacements when the shotcrete has fractured it still has significant retaining capabilities.
The shotcrete also prevents corrosion of the mesh and it strengthens the rock mass by
suppressing dilation.
The energy absorption capacity is another important property and differs between support
elements. Regular bolts can dissipate 1 – 5 kJ, while yielding bolts can dissipate about 30 kJ
(Kaiser et al., 1995). The energy absorption of mesh or shotcrete depends on the deformed
area. If the deformation is smaller than 200 mm the energy dissipation of mesh alone is quite
low, so the holding elements would have to take care of most of the energy. In combination
91
with shotcrete, however, the capacity is 3-5 times higher than for mesh alone, for
displacements of 50-100mm.
Kaiser et al. (1995) summarized a number of tests performed and formed design values for
load, displacement and energy absorption capacities for some common support elements used
in Canada, see Table 7-2.
Table 7-2. Design values for load-displacement parameters of support elements, after
Kaiser et al. (1995).
The energy absorption of the shotcrete and welded-wire mesh combination (marked with * in
Table 7-2) is valid at displacements below 100 to 150 mm. The displacement limit and energy
absorption in Table 7-2 were taken at the point of failure for rock bolts and at the peak load
for mesh and shotcrete (Kaiser et al., 1995). The gauge numbers of the mesh corresponds to
the diameter of the wire (www.screentg.com). A 4 gauge wire has a diameter of 5.723 mm, a
6 gauge wire has a diameter of 4.877 mm, and a 9 gauge has a diameter of 3.767 mm
(www.twpinc.com).
92
have to be well connected to each other to make the system effective. Kaiser et al. (1995)
noted three common weaknesses of support systems.
- the use of small rockbolt plates to connect to mesh and retaining elements. A strong
connection can be formed using larger rockbolt plates or load spreaders.
- too small overlap for mesh sheets can form a line of weakness. This can be avoided using an
overlap of at least one mesh square, use of mesh clips or addition of shotcrete.
- lacing can transfer significant loads to the holding elements, so the connections are critical.
It is important that no bending moments or stress concentrations occur at these connections.
In order to have an efficient support system it is essential that the required properties are
matched to the anticipated rockburst damage. Sometimes the rockbursts can be so severe that
they cause violent ejection of rock despite a reasonable support system. This upper limit of
energy absorption is called Maximum Practical Support Limit (MPSL) and lies around a
capacity of 50 kJ/m2 (Kaiser et al., 1995). When this limit is reached other measures have to
be considered, like changing excavation geometry, mine sequencing or destressing. Several
factors contribute to the MPSL, but it generally occurs when the thickness of the fracture zone
around the excavation exceeds 1.5 – 2 m, or when wall displacement exceeds 300 mm (Kaiser
et al., 1995).
93
94
8 CONCLUDING REMARKS
This literature review is not intended to cover every aspect of seismicity, but instead to cover
basic concepts and terminology. The literature review will also provide the background
information for selection of important parameters, to be studied in the cases included in the
Licentiate Thesis.
The importance of the in-situ stress state in the rock mass around the mine cannot be
underestimated. Knowledge of how mining influences this stress state is important; – back-
analysis of failures can provide valuable information. Mining should be planned keeping
seismicity in mind, since a small change in layout or sequence can improve stress conditions
around the stopes. The size of stopes, and the amount of rock blasted in each round affect the
amount of available seismic energy. Incremental mining reduces the amount of energy
available as seismic waves, but it can be favorable to blast a large round, because much of the
seismicity occurs directly after blasting. The reason for this is that the blast gives the rock
mass a “shake”, so that structures and rock close to failure actually fail. For a large blast no
personnel are allowed underground, so by the time all the gases have been ventilated most of
the stress redistribution has already taken place. Thus, the risk of a seismic event that can
endanger personnel in the vicinity of the stope has been greatly reduced. A system for
mapping of seismic failures has to be developed, to ensure that seismically active areas are
known, so that e.g., the correct reinforcement is used. In mines where seismicity is a problem,
and the mining method allows, backfill should be considered. Backfill has many advantages;
it provides support for stope walls – preventing progressive spalling, it prevents closure of
stopes, and it absorbs energy released by seismic events.
Another important factor to be considered is the location of geologic structures both in the
mine and in the surrounding rock mass. As mining progresses the faults may become
seismically active, and the consequences of slip can be disastrous. When planning the
production, mining should start at the fault and proceed away from it, where possible. If this
mining sequence is impossible, the support has to be designed with dynamic behavior in
mind.
Seismic monitoring should be considered in mines that experience seismicity, since it not only
can be used to locate and determine the size of seismic events, but also can provide valuable
95
information about the state of the rock mass surrounding the mine. Using data from the
seismic system for calibration of numerical models is quite common in mechanized hard rock
mines (Potvin and Hudyma, 2001). Analyzing the seismic source parameters can be very
valuable, but requires a substantial effort and time. Large volumes of data are needed, and the
variations in the seismic source parameters must be studied in localized areas, by isolating
individual sources and source mechanisms. An example is the study by Alcott et al. (1998),
where temporal increases in seismic moment, seismic energy, and apparent stress were
correlated with major failures occurring in the mine. The data used for analysis was hand-
picked, which ensures high quality data, but requires expertise and time. Usually there is only
one person in charge of the seismic system, which is usually an engineer with limited time to
spend on analysis. To perform this kind of seismological analysis routinely, better automatic
picking of first arrivals by the system is necessary, along with development of automatic tools
to give a first estimate.
The conclusion from this literature review is that the parameters that should be studied in the
mines are:
Destressing, reinforcement practices, and the use of data from seismic monitoring systems
should also be studied. This information can provide ideas for the Swedish mines starting to
experience seismicity or that have decided to invest in seismic monitoring systems.
96
9 REFERENCES
Alcott, J. M., Kaiser, P. K., Simser, B. P., 1998, Use of Microseismic Source Parameters for
Rockburst Hazard Assessment. Pure Appl. Geophys., vol 153, pp41-65
Amadei, B., Pan, E., 1992, Gravitational stresses in anisotropic rock masses with inclined
strata. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., vol 29, pp 225-236
Amadei, B., Savage, W. Z., Swolfs, H. S., 1987, Gravitational stresses in anisotropic rock
masses. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., vol 24, pp 5-14
Amadei, B., Stephansson, O., 1997, Rock Stress and Its Measurement. London: Chapman and
Hall
Andersen, L., 2001, A relative moment tensor inversion technique applied to seismicity
induced by mining. Ph.D. Thesis, University of Witwatersrand, Johannesburg, South Africa
Arabasz, W. J., Nava, S. J., Phelps, W. T., 1997, Mining seismicity in the Wasatch Plateau
and Book Cliffs coal mining districts, Utah, USA. Proc. 4th Int Symp Rockbursts and
Seismicity in Mines / Kraków/ 11-14 August 1997, Gibowicz and Lasocki (eds.), Rotterdam:
A.A. Balkema, pp 111-116
Arvidsson, R., Bodare, A., Kulhánek, O., Persson, L., 2000, Essential Seismology. Lecture
notes, Department of Earth Sciences, Uppsala University.
van Aswegen, G., Mendecki, A. J., Funk, C., 1997, Chapter 11 in Seismic Monitoring in
Mines, Mendecki (ed.), London: Chapman & Hall
Beck, D., 2000, A method for Engineering Management of Induced Seismicity in Deep-Level
Hard Rock Mining. Ph.D. Thesis, University of Queensland, Australia (unpublished)
Beck, D. A., Brady, B. G. H., 2002, Evaluation and application of controlling parameters for
seismic events in hard-rock mines. Int. J. Rock Mech. Min. Sci., vol 39, pp 633-642
Bigarre, P., Ben Slimane, K., Tinucci, J., 1993, 3-Dimensional modelling of fault-slip
rockbursting. Proc. 3rd Int Symp Rockbursts and Seismicity in Mines / Kingston/ 16-18 August
1993, Young (ed.), Rotterdam: A.A. Balkema, pp 315-319
97
Board, M., Rorke, T., Williams, G., Gay, N., 1992, Fluid injection for rockburst control in
deep mining. Proc. 33rd US Symp Rock Mechanics, / Santa Fe N.M./ 3-5 June 1992, Tillerson
and Wawersik (eds.), Rotterdam: A.A. Balkema, pp 111-120
Bolt, B. A., 1993, Earthquakes and Geological Discovery. New York: Scientific American
Library, USA. ISBN: 0-7167-5040-6
Brummer, R. K., Rorke, A. J., 1990, Case studies on large rockbursts in South African Gold
mines. Proc. 2nd Int Symp Rockbursts and Seismicity in Mines / Minneapolis/ 8-10 June 1988,
Fairhurst (ed.), Rotterdam: A.A. Balkema, pp 323-329
Brune, J. N., 1979, Implications of earthquake triggering and rupture propagation for
earhquake preidction based on premonitory phenomena. Journal of Geophysical Research,
vol 84, pp 2195-2198
Cook, N. G. W., 1976, Seismicity induced by mining. Engineering Geology, vol 10, pp. 99-
122
Gaviglio, P., Revalor, R., Piguet, J. P., Dejean, M., 1990, Tectonic structures, strata properties
and rockbursts occurrence in French coal mine. Proc. 2nd Int Symp Rockbursts and Seismicity
in Mines / Minneapolis/ 8-10 June 1988, Fairhurst (ed.), Rotterdam: A.A. Balkema, pp 289-
293
Gibowicz, S. J., Kijko, A., 1994, An Introduction to Mining Seismology. San Diego:
Academic Press, ISBN: 0-12-282120-3
Gill, D. E., Aubertin, M., Simon, R., 1993, A practical engineering approach to the evaluation
of rockburst potential. Proc. 3rd Int Symp Rockbursts and Seismicity in Mines / Kingston/ 16-
18 August 1993, Young (ed.), Rotterdam: A.A. Balkema, pp 63-68
Hanks, T. C., Kanamori, H., 1979, A Moment Magnitude Scale. J. Geophys. Res., vol 84, pp
2348-2350
98
Hedley, D. G. F., 1992, Rockburst Handbook for Ontario Hardrock Mines. CANMET Special
Report SP92-1E, Ottawa: Canada Communication Group
Hoek, E., Brown, E. T., 1982, Underground Excavations in Rock 2nd ed. Institution of Mining
and Metallurgy, London: E and FN Spon ISBN: 0-419-16030-2
Hudson, J. A., Cooling, C. M., 1988, In situ rock stresses and their measurement in the UK –
Part I. The current state of knowledge. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., vol
25, pp 363-370
Hyett, A. J., Dyke, C. G., Hudson, J. A., 1986, A critical examination of basic concepts
associated with the existence and measurement of in-situ stress. In Proc. Int. Symp. on Rock
Stress and Rock Stress Measurements / Stockholm, Luleå: Centek Publ. , pp 387-391
Jaeger, J.C., Cook, N. G. W., 1979, Fundamentals of Rock Mechanics 3rd ed. London:
Chapman and Hall, ISBN: 0-412-22010-5
Jeremic, M. L., 1987, Ground mechanics in hard rock mining. Rotterdam: A.A. Balkema,
ISBN: 90-6191-587-2
Johnston, J. C., Einstein, M. H., 1990, A survey of mining associated seismicity. Proc. 2nd Int
Symp Rockbursts and Seismicity in Mines / Minneapolis/ 8-10 June 1988, Fairhurst (ed.),
Rotterdam: A.A. Balkema, pp 121-127
Kaiser, P. K., Maloney, S. M., 1997, Ground motion parameters for design of support in
burst-prone ground. Proc. 4th Int Symp Rockbursts and Seismicity in Mines / Kraków/ 11-14
August 1997, Gibowicz and Lasocki (eds.), Rotterdam: A.A. Balkema, pp 337-342
Kaiser, P. K., McCreath, D. R., Tannant, D. D., 1995, Rockburst Support, Volume 2 of
Canadian Rockburst Research Program 1990-1995, vols 1 – 6. CAMIRO Mining Division,
Sudbury, Ontario, Canada
99
Kleczek, Z., Zorychta, A., 1993, Coal bumps induced by mining tremors. Proc. 3rd Int Symp
Rockbursts and Seismicity in Mines / Kingston/ 16-18 August 1993, Young (ed.), Rotterdam:
A.A. Balkema, pp 87-94
Knoll, P., Kuhnt, W., 1990, Investigation of the mechanism of rockbursts by seismological
measurements. Rockbursts, Global Experiences, International Bureau of Strata Mechanics,
Ghose, and Rao (eds.), Rotterdam: A.A. Balkema, pp. 11-36
Krishnamurthy, R., Shringarputale, S. B., 1990, Rockburst hazard in Kolar Gold Fields. Proc.
2nd Int Symp Rockbursts and Seismicity in Mines / Minneapolis/ 8-10 June 1988, Fairhurst
(ed.), Rotterdam: A.A. Balkema, pp 411-420
McGarr, A., 1984, Some applications of seismic Source Mechanism Studies to Assessing
Underground Hazard. Proc. 1st Int Symp Rockbursts and Seismicity in Mines / Johannesburg/
1982, Gay and Wainwright (eds.), Rotterdam: A.A. Balkema, pp 199-208
Mendecki, A. J., van Aswegen, G., Mountfort, P., 1999, A guide to routine seismic
monitoring in mines, in A Handbook on Rock Engineering Practice for Tabular Hard Rock
Mines. Jager and Ryder (eds.) SiMRAC, Cape Town, South Africa: Creda Communications,
Morrison, D. M., MacDonald, P., 1990, Rockbursts at INCO Mines. Proc. 2nd Int Symp
Rockbursts and Seismicity in Mines / Minneapolis/ 8-10 June 1988, Fairhurst (ed.),
Rotterdam: A.A. Balkema, pp 263-267
Mountfort, P., Mendecki, A. J., 1997, Chapter 1 in Seismic monitoring in mines. Mendecki
(ed.), London: Chapman & Hall
Müller, B., Zoback, M. L., Fuchs, K., Mastin, L., Gregersen, S., Pavoni, N., Stephansson, O.,
Ljunggren, C., 1992, Regional patterns of tectonic stress in Europe. J. Geophys. Res., vol 97,
11783-803
100
Myrvang, A. M., Alnæs, L., Hansen, S. E., Davik K. I., 1997, Heavy spalling problems in
road tunnels in Norway – long time stability and performance of sprayed concrete as rock
support. Proc. Int. Symp. Rock Support, Lillehammer 23 – 25 June 1997. Rotterdam: A.A.
Balkema
Ortlepp, W. D., Stacey, T. R., 1994, Rockburst Mechanisms in Tunnels and Shafts.
Tunnelling and Underground Space Technology, vol 9:1, pp 59-65
Ortlepp, W. D., 1997, Rock Fracture and Rockbursts, an illustrative study. Johannesburg: The
South African Institute of Mining and Metallurgy
Petukhov, I. M., 1990, Theory of rockbursts and their classification. Rockbursts, Global
Experiences, International Bureau of Strata Mechanics, Ghose, and Rao (eds.), Rotterdam:
A.A. Balkema, pp 3-9
Poplawski, R. F., 1997b, Seismic parameters and rockburst hazard at Mt Charlotte Mine, Int.
J. Rock Mech. Min. Sci., vol 34:8, pp 1213-1228
Potvin, Y., Hudyma, M. R., 2001, Keynote address: Seismic monitoring in highly mechanized
hardrock mines in Canada and Australia. Proc. 5th Int Symp Rockbursts and Seismicity in
Mines / Johannesburg/ 17-20 September 2001, van Aswegen, Durrheim and Ortlepp (eds.),
Johannesburg: South African Institute of Mining and Metallurgy, pp 267-280
Reinecker, J., Heidbach, O., Tingay, M., Connolly, P., Müller, B., 2004, The 2004 release of
the World Stress Map. (available online at www.world-stress-map.org)
Ryder, J. A., 1988, Excess shear stress in the assessment of geologically hazardous situations.
Journal of the South African Institute of Mining and Metallurgy, vol 88:1 pp 27-39
Salamon, M. D. G., 1983, Rockburst hazard and the fight for its alleviation in South African
gold mines. Proc. Symp. Rockbursts: prediction and control / London / 20 October 1983,
London: Institution of Mining and Metallurgy, pp 11-36
101
Scott, D. F., 1990, Relationship of geologic features to seismic events, Lucky Friday Mine,
Mullan, Idaho, Proc. 2nd Int Symp Rockbursts and Seismicity in Mines / Minneapolis/ 8-10
June 1988, Fairhurst (ed.), Rotterdam: A.A. Balkema, pp 401-405
Scott, D. F., Williams, T. J., Friedel, M. J., 1997, Investigation of a rock-burst site, Sunshine
Mine, Kellogg, Idaho. Proc. 4th Int Symp Rockbursts and Seismicity in Mines / Kraków/ 11-14
August 1997, Gibowicz and Lasocki (eds.), Rotterdam: A.A. Balkema, pp 311-315
Sjöberg, J., Lundman, P., Nordlund, E., 2001, Analys och prognos av utfall i bergschakt, KUJ
1045 Slutrapport, LKAB internal report nr 01-762 (In Swedish)
Spottiswoode, S. M., 1990, Volume excess shear stress and cumulative seismic moments.
Proc. 2nd Int Symp Rockbursts and Seismicity in Mines / Minneapolis/ 8-10 June 1988,
Fairhurst (ed.), Rotterdam: A.A. Balkema, pp 39-43
Toper, A. Z., Grodner, M., Stewart, R. D., Lightfoot, N., 1997, Preconditioning: A rockburst
control technique. Proc. 4th Int Symp Rockbursts and Seismicity in Mines/ Krákow/ 11-14
August 1997, Gibowicz and Lasocki (eds.), Rotterdam: A.A. Balkema, pp 267-272
Toper, A. Z., Stewart, R. D., Kullmann, D. H., Grodner, M., Lightfoot, N., Janse van
Rensburg, A. L., Longmore, P. J., 1998, Develop and implement preconditioning techniques
to control face ejection rockbursts for safer mining in seismically hazardous areas. Draft of
final project report project number GAP 336, Safety in Mines Research Advisory Committee,
Rock Engineering Programme, CSIR Division of Mining Technology
Vervoort, A., Moyson, D., 1997, Steel fibre reinforced shotcrete: An adequate support for
rockburst conditions? Proc. 4th Int Symp Rockbursts and Seismicity in Mines / Kraków/ 11-14
August 1997, Gibowicz and Lasocki (eds.), Rotterdam: A.A. Balkema, pp 355-359
102
Warpinski, N. R., Teufel, L. W., 1991, In-situ stress measurements at Rainier Mesa, Nevada
Test Site – influence of topography and lithology on the stress state in tuff. Int. J. Rock Mech.
Min. Sci. & Geomech. Abstr., vol 28, pp 143-161
Wawersik, W. R., Fairhurst, C., 1970, A study of brittle rock fracture in laboratory
compression experiments. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., vol 7, pp 561-575
Wiles, T. D., 1998, Correlation Between Local Energy Release Density and Observed
Bursting Conditions at Creighton Mine. Report under Contract for INCO Ltd. Mines
Research, Sudbury, Canada
Wong, I. G., 1992, Recent developments in rockbursts and mine seismicity research. Proc.
33rd US Symposium on Rock Mechanics / Santa Fe, New Mexico/ 3-5 June 1992, Tillerson
and Wawersik (eds.), Rotterdam: A.A. Balkema, pp 1103-1112
Yi, X., Kaiser, P., 1993, Mechanism of rock mass failure and prevention strategies in
rockburst conditions. Proc. 3rd Int Symp Rockbursts and Seismicity in Mines / Kingston/ 16-18
August 1993, Young (ed.), Rotterdam: A.A. Balkema, pp 141-145
103