You are on page 1of 67

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/303347923

Applied Mine Seismology Concepts and Techniques

Technical Report · January 2010


DOI: 10.13140/RG.2.1.2781.0162

CITATION READS
1 834

1 author:

Marty Hudyma
Laurentian University
66 PUBLICATIONS 424 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Data Driven Seismic Hazard Evaluation in Canadian Underground Mines View project

Seismic monitoring in caving mines View project

All content following this page was uploaded by Marty Hudyma on 19 May 2016.

The user has requested enhancement of the downloaded file.


APPLIED
MINE SEISMOLOGY
CONCEPTS AND
TECHNIQUES

Technical Notes for ENGR 5356 – Mine Seismic Monitoring Systems

Dr. Marty Hudyma, Assistant Professor


Laurentian University
January 2010
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES


The following document discusses the background of some of the more commonly used
techniques in mine seismology, relating seismic events to rockmass failure in mines.

Contents
1 Introduction .......................................................................................................................................................1
1.1 Independent Seismic Source Parameters ...................................................................................................1
1.1.1 Event Time..........................................................................................................................................1
1.1.2 Event Location....................................................................................................................................3
1.1.3 Seismic Energy ...................................................................................................................................5
1.1.4 Seismic Moment .................................................................................................................................5
1.1.5 Source Size .........................................................................................................................................6
1.2 Seismic Hazard ..........................................................................................................................................7
1.3 Seismic Source Mechanism .....................................................................................................................10
1.3.1 Typical Seismic Source Mechanisms in Mines ................................................................................10
1.3.2 Determining Seismic Source Mechanism.........................................................................................13
1.3.2.1 Waveform Techniques...............................................................................................................13
1.3.2.2 Issues in Applying Direct Waveform Techniques for Determining Source Mechanism ..........15
1.3.2.3 Inferred Seismic Source Mechanism Techniques......................................................................15
1.3.2.4 Issues in Applying Indirect Waveform Techniques for Determining Source Mechanism........16
1.4 Important Conditions for Successful Seismic Data Analysis ..................................................................17
1.4.1 Seismic System Limitations..............................................................................................................17
1.4.2 Practical Considerations ...................................................................................................................19
2 Seismic Hazard................................................................................................................................................21
2.1 Event Magnitude......................................................................................................................................21
2.1.1 Description........................................................................................................................................21
2.1.2 Limitations in Magnitude Estimation ...............................................................................................22
2.2 Frequency-Magnitude Analysis ...............................................................................................................24
2.3 Magnitude Time History Analysis...........................................................................................................27
2.3.1 Introduction.......................................................................................................................................27
2.3.2 Other Considerations ........................................................................................................................30
2.4 Instability Analysis ..................................................................................................................................33
2.5 Apparent Stress Time History..................................................................................................................35
2.6 Daily Histogram of Event Frequency ......................................................................................................40
3 Seismic Source Mechanism ............................................................................................................................41
3.1 S:P Energy Ratio Analysis.......................................................................................................................41
3.1.1 Background.......................................................................................................................................41
3.1.2 Additional Considerations ................................................................................................................44
3.2 Diurnal Analysis ......................................................................................................................................47
3.3 Phasor Analysis........................................................................................................................................49
4 Data Integrity ..................................................................................................................................................51
4.1 Daily Histogram.......................................................................................................................................51
4.2 Magnitude-Time History..........................................................................................................................52
4.3 Energy-Moment Relation.........................................................................................................................55
4.4 Gutenberg-Richter Frequency-Magnitude Relation.................................................................................57
References ..............................................................................................................................................................61

Marty Hudyma
Assistant Professor, Laurentian University i
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

1 INTRODUCTION

Effective seismic analysis involves identifying trends in seismic data and relating those trends to typical
rockmass failure mechanisms using populations of seismic events. These trends also give information about
seismic hazard (or the largest expected seismic event) and seismic source mechanism (the rockmass failure mode
which cause the seismic events).

1.1 Independent Seismic Source Parameters

To describe a seismic event quantitatively, Mendecki et al. (1999) suggested the following information is
required:
 location of the event,
 time of the event, and
 at least two independent seismic source parameters, such as seismic energy, seismic moment, or
source size.

These are independent seismic source parameters that quantitatively describe a seismic event and can be used to
characterise seismic events or a population of events. The following section describes each of the five
independent seismic parameters.

1.1.1 Event Time

The time of a seismic event identifies the occurrence of the seismic-related rockmass failure. Event time
compared to recent mine blasting can be an important consideration. The smaller the time interval blasts and
seismic events, potentially the greater the relation between the mine blast and the seismic-related rockmass
failure.

Often it is the time relative to other seismic events that is of most interest. A commonly used measure in seismic
hazard characterisation is the frequency of occurrence of events of a given size, expressed as:

λ=1/t
where,
λ = the frequency of occurrence
t = the time between events of a certain magnitude
This term is widely used in earthquake seismology to investigate the expected return time of an earthquake of a
given size.

In mines, the time of occurrence of most events is directly related to mine blasting. Seismicity is fundamentally
induced by stress changes due to blasting-induced stope geometry changes. Figure 1 shows a histogram of the
number of events per day in a Western Australian mine. The peaks in event rate are strongly correlated to stope
blasts in the mine.

Marty Hudyma
Assistant Professor, Laurentian University 1
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 1 Histogram of the number of events per day in a Western Australian mine (Hudyma et al.,
2003). Increases in the number of events per day are strongly correlated to the occurrence
of stope blasting in the mine.
While time is conceptually the simplest seismic source parameter to measure and comprehend, it can be one of
the most perplexing and difficult parameters to utilise and understand in mining-induced seismicity. Brummer
(1999) reports that the larger the event, the more random the pattern in time of occurrence. This may signify that
the rockmass failure processes for very large seismic events are fundamentally driven by factors different than
smaller seismic events, or that the root rockmass failure mechanisms are much broader than for smaller events.
For instance, small events may be directly induced by a discrete mine blast, but very large events may be as a
result of many years of extraction in an orebody.

Basson and Ras (2005) investigated time-space relations between events at Target gold mine. They found bigger
events were uncorrelated in time with smaller events, but were better correlated with other bigger events in the
vicinity in time. They also found dramatically different time-space correlations for the different faults and dykes
in the mine.

Simple cause and effect relations in time and space frequently often cannot explain the occurrence of seismic
events. Urbancic and Trifu (1995) describe an episode of seismicity at Strathcona mine. A mine blast triggered
a large seismic event in the mine. Over the next four days, nine more large seismic events occurred in mine
pillars up to 200 m from the mine blast and initial seismic event. A complex combination of stress migration
and the influence of mine faults were considered the main reasons for the episode of large seismic events.
Similar interactions between relatively distant sources of seismicity are reported in literature (Kijko et al., 1993;
Hudyma, 1995).

It is commonly found that the seismic response to a particular mine blast can be disproportionate to the size of
the blast (Urbancic and Trifu, 1995; Hudyma, 1995). Morrison et al. (1993) report similar results with
seismicity occurring in distinct episodes despite relatively uniform mine production rates. They suggest that the
naturally complex rockmass and mining environment results in chaotic behaviour of mine seismicity, in which
the cause of smaller events are intrinsically simple, but the causes of large events are extremely complex.

The timing of seismicity is crucial to effective management of safety risks associated with mine seismicity. If
large events follow closely after significant mine blasts, then creating temporary exclusion zones in the

Marty Hudyma
Assistant Professor, Laurentian University 2
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

proximity of the blast is a very effective means of mitigating the safety risk of large seismic events. However, if
large seismic events occur unrelated to mine blasts, temporary exclusion ceases to be an effective risk mitigation
technique as mine operations can rarely be interrupted indefinitely in anticipation of a large damaging seismic
event.

1.1.2 Event Location

For mining-induced seismicity, accurate locations is the most important seismic source parameter (Brummer,
1999). Accurate spatial location of seismic data on mine plans frequently identifies the location of seismic
source near geological structures, or mine structures such as stopes and pillars (Leslie and Vezina, 2001;
Hudyma et al., 1995; Amidzic, 2001). For example, Figure 2 shows seismic events occurring along a fault in the
footwall of an orebody (Mollison et al., 2003). The location of the seismic events identify the location, as well
as strike and dip of the fault.

Figure 2 Seismic events show the location, strike and dip of a fault (Mollison et al., 2003).

Seismicity in a mine may look hypothetically similar to the diagrammatic sketch in Figure 3. The different
symbols represent events of different magnitudes. There may be several seismic sources in a mine creating
seismic events concurrently as suggested in Figure 4. Importantly, the seismicity at each seismic source is a
result of the local failure mechanism, with each seismic source potentially having a different seismic hazard
potential.

Marty Hudyma
Assistant Professor, Laurentian University 3
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 3 Seismic events around a proto-typical undergound open stope mine (Hudyma et al., 2003).

Figure 4 Each group of seismic events represents a separate seismic source, with each seismic
source generating seismicity concurrently (Hudyma et al., 2003).

Marty Hudyma
Assistant Professor, Laurentian University 4
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

1.1.3 Seismic Energy

Radiated seismic energy is the energy released at the source of seismic event that is radiated as seismic waves.
According to Gibowicz and Kijko (1994), the radiated seismic energy can be calculated for a P-wave or an S-
wave as:

Jc
E  4   c R  
Fc 2

where,
E = radiated energy (joules)
ρ = rock density (kg/m3)
c = velocity of the wave in rock (m/s)
R = the distance from the seismic source (m)
Jc = the integral of the square of the ground velocity

1.1.4 Seismic Moment

Seismic moment is widely accepted as the best means of estimating the size of a slip-related seismic event.
From seismic waveform data, Gibowicz and Kijko (1994) note that the seismic moment can be derived with the
following formula:

o
Mo  4  c  R
Fc

where,
Mo = seismic moment (Nm)
ρ = rock density (kg/m3)
c = velocity of the wave in rock (m/s)
R = the distance from the seismic source (m)
Ωo = the low frequency plateau of the frequency spectrum of a seismic waveform (Figure 5)
Fc = an empirical radiation pattern coefficient

Marty Hudyma
Assistant Professor, Laurentian University 5
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 5 Low frequency plateau (Ωo) and corner frequency (fo) for a typical seismic event (Hedley,
1992).

1.1.5 Source Size

Source size is a model dependant parameter describing the dimension of the seismic event failure surface.
Source size is commonly defined as a radius or diameter using:

l = c / fo

where,
l = source size (m)
c = model dependent constants
fo = corner frequency, in the frequency domain, the intersection of the low frequency plateau with
the high frequency spectrum decay (Figure 5)

Marty Hudyma
Assistant Professor, Laurentian University 6
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

1.2 Seismic Hazard

Understanding seismic hazard in mines is one of the fundamental objectives in seismic monitoring in mines.
Seismic hazard is commonly evaluated as a measure of the largest event that may occur. This is important as it
dictates the level of strong ground motion that may be induced by a seismic event, which is closely related to the
potential for rockmass damage. A commonly referenced definition of seismic hazard is:

“An estimation of the mean probability (over space and time) of the occurrence of a seismic
event with a certain magnitude within a given time interval.” Gibowicz and Kijko (1994)

The challenges in estimating seismic hazard are clearly highlighted in this definition with regard to the
uncertainty involved in “mean probability”, “certain magnitude” and “within a give time interval”.

In mines, some important considerations in seismic hazard assessment include:


 Seismic hazard varies over time. In many mines, blasting related stress change is a dominating, but
discontinuous influence on the occurrence of seismicity in mines. The likelihood of potentially
hazardous seismicity is elevated during periods of stope blasting.
 Seismic hazard varies in space. Large seismic events are a result of large-scale rockmass failures,
which typically occur in areas of high induced stress and high levels of mine extraction. In many
cases, the vast majority of mine development will not suffer any significant rockmass failure and
will be relatively aseismic.

For example, Figure 6 shows 367 events larger than magnitude 0 at Mount Charlotte mine in Australia, over a
time period of 5 years (Hudyma, 2008). Seismic hazard varies in space and time, with the timing of events
generally following stope blasts in the mine. Some key observations from Figure 6 include:
 The location and number of the seismic events varies considerably from one year to the next.
 On a yearly basis, the seismic events show distinct spatial clusters, largely associated with the flat
dipping faults in the mine.
 The number of events generally increases with depth.
 Over the full 5-year time period, there are areas of the mine with large numbers of seismic events,
and areas of the mine with virtually no large seismic events.

Marty Hudyma
Assistant Professor, Laurentian University 7
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 6 Seismic events greater than Richter magnitude 0 at Mount Charlotte Mine are shown on a
yearly basis (top left – 1994, top right – 1995, middle left – 1996, middle right – 1997,
bottom left – 1998, bottom right – 1994 to 1998). The charts demonstrate the spatial and
temporal variations of event occurrence and seismic hazard in the mine (redrawn from
Hudyma, 2008).

Marty Hudyma
Assistant Professor, Laurentian University 8
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

There are a number of techniques that may be used to evaluate seismic hazard. van Aswegen (2005) writes that
the techniques commonly used depend on the time frame of interest. He suggests that seismic hazard in mines
are commonly evaluated over three time frames: long-term seismic hazard, medium term seismic hazard and
short-term seismic hazard.

 Long-term seismic hazard is controlled by mine-wide geological features and mining parameters.
Long-term seismic hazard generally considers the statistical likelihood of occurrence of an event of
a give size, or the expected return period for an event of a certain size (Cornell, 1968; Kijko, 1997).
van Aswegen (2005) notes that long-term seismic hazard is generally evaluated using numerical
models calibrated to the past seismic response to mining.
 Medium-term seismic hazard is evaluated on about a monthly basis, using trends in seismic source
parameters to identify the potential for areas prone to large events (Van Aswegen, 2005). A key
question is how well does medium-term seismic hazard, derived from past seismic events, identify
future high seismic hazard?
 Short-term seismic hazard evaluation endeavours to identify unstable rockmass conditions
potentially leading to large, damage-causing seismic events. In contrast, short-term seismic hazard
is commonly used to evaluate the potential occurrence of seismic events in the next hours or few
days (van Aswegen, 2005). This is analogous to short-term prediction of seismic events. A number
of parametric and non-parametric methods have been proposed for evaluating short-term seismic
hazard (van Aswegen and Butler, 1993; Mendecki et al., 1999). Figure 7 is a schematic illustration
of the use of a parametric to evaluate short-term seismic hazard, or the probability of a seismic event
of a certain size (Kijko, 1997).

Figure 7 Schematic illustration of the use of a parametric approach for seismic event prediction
(Kijko, 1997). A parametric technique is used over a time window (t – Δt’) to estimate the
probability of an event of a particular size in the time window (t + Δt).

Marty Hudyma
Assistant Professor, Laurentian University 9
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

1.3 Seismic Source Mechanism

1.3.1 Typical Seismic Source Mechanisms in Mines

Gibowicz (1990) states that “two broad types of mine tremors are almost universally observed: (a) those directly
connected with mining operations, i.e., those associated with the formation of fractures at stope faces, and (b)
those that are not, i.e., those associated with movement on major geological discontinuities.” While this
categorisation is more related to the general influences causing the seismicity rather than the explicit rockmass
failure mechanism causing the events, it does make a key distinction between seismic events that are directly
induced by mining versus those events that are triggered by mining. This distinction is important, as events
induced by mining are largely reliant on energy changes in the mining to cause rockmass failure. In this case,
the seismic response to mining will be relatively proportional to the scale and influence of the mining. However,
for triggered seismicity, the rockmass may already be in a state of instability, requiring a relatively small energy
change from mining to cause large scale failure and a large seismic event.

Hasegawa et al. (1989) propose six basic mechanisms of mining-induced tremors (Figure 8), including cavity
collapse, pillar burst, tensional fault, and three types of fault-slip. The far field seismic energy radiation pattern
basic for each mechanism is shown for each of the six basic mechanisms.

Figure 8 Six basic mechanisms of mining-induced tremors (Hasegawa et al., 1989).

Ortlepp (1992) discusses five mechanisms of damaging rockbursts (Table 1), listing the typical rockmass
movement, the direction of first motion that would be recorded by seismic sensors, and the approximate
maximum magnitude range for events of each mechanism.

Marty Hudyma
Assistant Professor, Laurentian University 10
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Table 1 Five mechanisms of damaging rockbursts proposed by Ortlepp (1992).

Seismic Event Postulated Source Mechanism First Motion from Richter


Seismic Records Magnitude
ML
Strain-burstingSuperficial spalling with violent Usually undetected, -0,2 to 0
ejection of fragments could be implosive
Buckling Outward expulsion of larger slabs Implosive 0 to 1,5
pre-existing parallel to opening
Pillar or face Sudden collapse of stope pillar, or Implosive 1,0 to 2,5
crush violent expulsion of rock from tunnel
face
Shear rupture Violent propagation of shear fracture Double-couple shear 2,0 to 3,5
through intact rockmass
Fault-slip Violent renewed movement on Double-couple shear 2,5 to 5,0
existing fault

These damage mechanisms relate to discrete, large-scale rockmass failures. Typically however, more than 90%
of seismic events recorded in mines are smaller than moment magnitude 0, representing smaller rockmass
failures than those described by Ortlepp and Hasegawa. A microseismic event rarely results in observable
rockmass damage, however, they may still provide valuable information about the larger damaging events.

Typical seismic source mechanisms of microseismic events in mines include all of the mentioned by Ortlepp
(1992), but would also include:
 Intact brittle rock fracture (Lynch et al., 2005),
 Coalescence of rockmass fractures such as rock joints (Trifu and Urbancic, 1996),
 High stresses acting in stope abutments (Andrieux and Simser, 2001),
 Crushing, shearing, and volumetric fracturing of mine pillars (Hedley, 1992), and
 Shear or rupturing of lithological contacts (Mollison et al., 2001).

Conceivably most rockmass failure mechanisms in brittle, hardrock have the potential to generate seismic
events. Figure 9 is a diagrammatic open stope mining block, with photographs of some of the commonly
observed rockmass failure mechanisms that generate seismic events. Some important aspects associated with
this diagram include:
 There are many seismic source mechanisms in mines.
 Different seismic source mechanisms may occur relatively close to each other.
 Seismic events are a result of local failure processes. A stope blast or large seismic event (in a
geometry such as in Figure 9), may induce seismic events to occur in each of the six different
seismic sources, however, the size and nature of the seismic events at each seismic source will be
predominantly related to the local rockmass failure mechanisms (fault movement, stress change,
stope overbreak, contrast in rockmass material properties and crushing of mine pillars).

Marty Hudyma
Assistant Professor, Laurentian University 11
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 9 Typical local rockmass failure mechanisms that potentially cause seismic events, (a) fault
movement, (b) stress change causing rockmass fracturing near excavations (c) stope
overbreak, (d) contrast in rockmass material properties causing strain-bursting, (e)
crushing of mine pillars, (f) stress increase causing rockmass deformation (after Hudyma,
2008).

Marty Hudyma
Assistant Professor, Laurentian University 12
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

1.3.2 Determining Seismic Source Mechanism

The techniques commonly used to determine seismic source mechanism can be divided into two methodologies.
Waveform techniques involve individual analysis of ground motion waveforms collected at a number of seismic
receivers for a particular event and determining information about the energy radiation pattern for that event. In
contrast, inferred techniques use empiricism and rules of thumb to suggest the seismic source mechanism,
generally looking at time and space trends in seismic source parameters of populations of tens and hundreds
seismic events.

1.3.2.1 Waveform Techniques

Moment Tensor Inversion

Moment tensors describe, in a first order approximation, the equivalent forces acting at seismic point sources.
The equivalent forces can be compared to source models, such as the displacement of a fault, sudden volume
decrease associated with implosional failure, and sudden volume increase due to explosional failure (Jost and
Herrmann, 1989). Moment tensor inversion is often deemed to be the best approach to study the failure mode of
mining-induced seismic events (Gibowicz, 1993), but it is also noted that there are few studies of moment
tensors for mining-induced seismicity. This is primarily owing to the need for exceptionally high quality seismic
data, requiring high quality digital waveforms from triaxial sensors and good focal coverage of the seismic
source. Data of this nature is rarely available in mines. In addition, the theoretical and mathematical complexity
of moment tensor analysis largely restricts its role to seismology research, rather than commonplace a minesite
tool.

First Motion Analysis

Numerous studies have found that for shearing along faults or pre-existing structural discontinuities, that the
double couple component of the moment tensor is an adequate representation of the forces acting at the seismic
source (Spottiswoode and McGarr 1975; Spottiswoode 1984; Gibowicz 1990). Urbancic (1991) and Duplancic
(2001) found that nodal plane orientations were similar to mapped fractures and known joint sets, suggesting
first motion analysis can be applied for minor rockmass discontinuities that have no history of shearing.

The double force couple is a component of the moment tensor, in which there are two opposing forces with no
net torque. The compressional wave first motion of ground displacement around a double-couple seismic source
is divided into four equal quadrants. Two of the quadrants have compressional first motions, and two of the
quadrants have dilatational first motions (Figure 10). Two nodal planes separate the four quadrants, with one of
the nodal planes corresponding to the plane of slip. The second nodal plane has no physical meaning.

Marty Hudyma
Assistant Professor, Laurentian University 13
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 10 The double couple force system, with a fault plane in the X2X3 plane and showing the two
quadrants of compressional (-) and two quadrants of dilatational (+) compressional wave
first motions (after Urbancic, 1991).

In first motion analysis, the direction of first motion and position of each sensor relative to the seismic source is
plotted on a stereographic projection. Typically, the two nodal planes are visually fit to the data (Figure 11).

Figure 11 Lower hemisphere projection of a fault plane solution (after Urbancic, 1991).
Compressional first motions are plotted as filled circles, dilatational first motions are
plotted as open circles, and low confidence first motion polarities are plotted as filled
triangles.

Marty Hudyma
Assistant Professor, Laurentian University 14
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

1.3.2.2 Issues in Applying Direct Waveform Techniques for Determining Source Mechanism

There are a number of problems and challenges in applying direct waveform techniques at minesites.
 Direct waveform techniques require high quality seismograms. Scattering, attenuation and noise in
the data caused by mine excavations and mine cultural noise (drills, fans and orepasses) often make
the majority of data collected unsuitable for direct waveform techniques.
 The seismic sensors need to surround a seismic source to adequately sample the radiation pattern of
energy from the seismic source. This is atypical of the seismic arrays found at most mining
operations. In most mines, the sensors tend to be on one side of the orebody, making seismic event
focal coverage inadequate. If there is insufficient sensor coverage around a seismic event, direct
waveform methods will likely give inconclusive results.
 For moment tensor analysis, at least six triaxial sensors are normally required. For most mines, it is
uncommon to have high quality seismic event waveforms on six triaxial sensors. In order to
routinely high quality seismic event waveforms of at least six triaxial sensors, an array of 8 to 10
triaxial sensors would likely be required.
 Conducting direct waveform analysis can be slow and painstaking work (Trifu and Shumila, 2002).
Consequently, mine site seismic system operators rarely use these techniques. When the techniques
are utilised in academic and research investigations, analysis is usually undertaken for relatively
small populations of events in a constrained time period or location.

These issues associated with direct waveform techniques often lead to ambiguous or inconclusive results for
seismic source mechanism determination. The likelihood of success of direct waveform techniques may be low
for events in some seismic systems in mines, depending on the number and type of sensors and the focal
coverage provided by the sensors. In a comprehensive first motion study at Ansil mine, McCreary et al. (1993)
found clear shearing or tensile seismic source mechanisms for less than 25% of the 382 events analysed. In the
study of a sequence of large seismic events using 22 triaxial sensors, Gibowicz (1997) found full moment tensor
solutions for 35 of the 199 events investigated. For a more typical mining seismic array of 8 triaxial sensors,
Trifu and Shumila (2002) undertook a moment tensor investigation at Kidd Creek mine. Although more than
1000 events per month are recorded in the mine, only between 1 and 4 events per month were retained for
moment tensor analysis.

1.3.2.3 Inferred Seismic Source Mechanism Techniques

There are a number of techniques that can be used to infer seismic source mechanism. These techniques tend to
be applied to a significant population of events. The premise is that the majority of seismic events within a
seismic source represent the same seismic source mechanism and will have similar trends in seismic source
parameters.

Seismic Event Source Location

Spatial plotting of high precision seismic data frequently identifies the location of seismogenic rock mass failure
near geological structures, or in proximity to mine blasts, mine stopes and mine pillars (Leslie and Vezina, 2001;
Hudyma et. al., 1995). Spatial plotting alone can frequently suggest the seismic source mechanism of mine
seismicity.

Marty Hudyma
Assistant Professor, Laurentian University 15
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

S:P Energy Ratio

As discussed earlier in this chapter, the ratio of S-wave to P-wave energy is strongly related to seismic source
mechanism. Fault-slip mechanism events commonly have S-wave to P-wave energy ratios in excess of 10
(Boatwright and Fletcher 1984, Cichowicz et al. 1990). While non-shear mechanisms such as stress fracturing,
tensile failure and volumetric stress change events commonly have S-wave to P-wave energy ratios in the range
of 1 to 3 (Urbancic et al., 1992). More information on S:P energy ratio analysis is found below in Section 3.1.

Frequency-Magnitude Relation

The slope of the Gutenberg-Richter frequent-magnitude relation can also be an indicator of seismic source
mechanism. Fault-slip related seismicity frequency has a very low b-value, often less than 0.8. While stress
change seismicity frequently has high b-value, often in the range of 1.2 to 1.5. More information on frequency-
magnitude analysis is found below in Section 2.2.

Diurnal Analysis

The time of day of occurrence of events can be an indicator of seismic source mechanism. Seismicity directly
associated stress with change due to mine blasting occurs directly after blasting. Seismicity associated with
geological structures often has little temporal relation to mine blasting. More information on diurnal analysis is
found below in Section 3.2.

Summary of Inferred Seismic Source Mechanism Techniques

The inferred seismic source mechanism techniques can be useful to suggest seismic source mechanism. In
contrast to the waveform methods of determining seismic source mechanism, they are easily applied to large
volumes of seismic data and can be readily applied to investigate temporal and spatial variations of seismic data.
Additional confidence about the seismic source mechanism can be inferred if more than one technique is used
for the same data set.

Importantly, inferred seismic source mechanism techniques require a significant population before they can be
used. If multiple mechanisms of events exist within the population of data, the techniques are likely to give an
ambiguous or indeterminate result.

1.3.2.4 Issues in Applying Indirect Waveform Techniques for Determining Source Mechanism

There are drawbacks and issues in determining seismic source mechanism using indirect methods:
 Indirect methods require that a significant population of data is available, and that the vast majority
of the events in the population are from the same seismic source. There is no 100% reliable means
of grouping seismic events by seismic source, although, spatial clustering can be used under some
circumstances.
 Indirect techniques often give ambiguous or inconclusive results if the population of events has
multiple mechanisms.

Marty Hudyma
Assistant Professor, Laurentian University 16
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

1.4 Important Conditions for Successful Seismic Data Analysis

A number of seismic source parameters are available to describe individual seismic events. However, it is a
central theme of this thesis that any single seismic event represents only a momentary component of a rockmass
failure process. In contrast, mine seismicity analysis techniques, incorporating hundreds or thousands of seismic
events, give a more holistic overview of a population of seismic events. These events may occur over a period
of months or years. Insight can be derived into the nature of a rockmass failure process, giving an understanding
of a number of valuable characteristics such as: space-time variations in the rockmass failure; changes in the rate
of rockmass failure; and the influence of mine blasting on rockmass failure.

However, there are some practical considerations and limitations that need to be regarded when applying seismic
source parameters and mine seismicity data analysis techniques to understand rockmass failure in mines.

1.4.1 Seismic System Limitations

Each individual seismic event represents a singular inelastic rockmass deformation at a moment of time during a
rockmass failure process. An individual seismic event may be only one of hundreds or thousands of recordable
events that emitted from a rockmass failure, occurring in a period of days, weeks or even months. So although
individual, larger events may be relatively important compared to individual, smaller events, it is the population
of events that occur over time that best describe the failure process.

It is also possible that any individual seismic event is not a component of the main failure process, but rather a
local rockmass response to a nearby failure process. For example, if a fault is slipping, not all of the events
recorded near the fault will be fault-slip events. Some of the events will be “secondary” rockmass re-
adjustments to the fault-slip on nearby geological discontinuities (such as joints, dykes and other conjugate
features). The seismic characteristics of these “secondary” events will be strongly related to the local rockmass
and stress conditions at the point of “secondary” re-adjustment, and may be dissimilar to the predominant fault-
slip failure mechanism. Often “secondary” seismic events occur in close proximity to the primary fault slip
failure. It may not be practical or possible to try to separate or clearly distinguish the secondary event
mechanisms from the primary event mechanisms. Ultimately, any seismic analysis technique will have to be
able to tolerate the occurrence of extraneous “secondary” seismic events.
The nature of seismic monitoring in mines is that not all events will be well collected. Seismic monitoring is a
sampling process, with practical limits of seismic monitoring system availability and data collection. Sometimes
the seismic system is not operational, or some of the sensors are not operational, and events are not captured. In
addition, a portion of the events will be poorly collected by seismic systems, with data being corrupted by
electrical noise or mechanical vibration, resulting in poor quality locations and unreliable seismic source
parameters. Events may occur simultaneously in different locations, making it difficult for the seismic system to
identify each event in the seismic record. Events, particularly large events, do occur during mine blasts and may
be poorly captured, or completely missed. Consequently, a seismic record will not be complete, but rather only a
sample of the seismic events that have occurred. A seismic data analysis technique needs to be functional using
only a reasonably representative sample of the total seismic record.

One of the potential solutions to understanding predominant failure mechanisms is to focus exclusively on the
largest events. There are a few benefits to this approach. Most of the smaller “secondary” events can be
removed from the analysis. The larger events are also usually better captured by seismic systems because they
are collected on many sensors. However, focussing only on the largest events has some potentially serious
pitfalls.
 The vast majority of the seismic record is not considered, so the majority of rockmass
deformation process is ignored.

Marty Hudyma
Assistant Professor, Laurentian University 17
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

 The issue of seismic system sampling is not alleviated. In fact, sampling issues are accentuated,
due to the dependence on collecting good quality seismic data for a fewer number of larger
events. Poorly capturing, or missing, key events may compromise the seismic data analysis.
 There may be less opportunity to use the seismic record to proactively characterise seismic
sources due to limited numbers of events. Smaller events, occurring early in the failure process,
may be excluded. Consequently, seismic analysis may become a reactive exercise, highlighting
only the most significant moments of a rockmass failure process.

A more reliable data analysis approach is to consider the entire population of events recorded. The larger events
are important, but only as highlights in the full seismic record. This approach is less dependent on the adequate
collection of the largest seismic events, and uses all of the available data to understand the seismic response to
mining. For example, Figure 12 shows two week of seismic events from a block caving mine (Hudyma, 2008).
Figure 12(a) is a frequency-magnitude relation for the data. Figure 12 (b) shows all events magnitude –2.0 and
larger, (c) magnitude –1.5, (d) magnitude ≥ –1.0, (e) magnitude ≥ –0.5, and (f) magnitude ≥ 0. There is far more
information in the image of 2258 events magnitude ≥ –2.0 than the eight events with magnitude ≥ 0.

Marty Hudyma
Assistant Professor, Laurentian University 18
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

b = -1.3 2258 events


2258

612

131

27

a b
612 events 131 events

c d
27 events 8 events

e f
Figure 12 Two weeks of seismic events recorded in a block caving mine (Hudyma, 2008).

1.4.2 Practical Considerations

There are some important conditions for seismic data analysis to be successful:
1. Analysis is much more effective if logical groups of seismic events are analysed together.
Typically, spatial clusters of events of similar seismic source mechanism are analysed together.
Alternatively, if data is taken from a very wide area (i.e. over an entire mine), or if the data

Marty Hudyma
Assistant Professor, Laurentian University 19
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

encompasses more than one failure mechanism, the results of the analysis are more likely to be
ambiguous or indeterminate.
2. There must be an adequate number of seismic events, over a reasonable time period, with no
extended periods of significant data loss. Typically an analysis group needs a few hundred to
several hundred seismic events over a time period of several months, before a level of confidence
can be associated with the analysis.
3. The analysis should not be reliant upon one or a couple of large seismic events. Seismic monitoring
is a means to collect the majority of seismic events that occur. Large and important seismic events
do occur during blasts and are frequently not captured or are poorly captured. Any analysis that is
highly dependent upon collecting and interpreting all of the large seismic events is not going to give
consistent results. Overall, it is better to use the population of seismic events to interpret the seismic
response to mining. It should be inconsequential if a few large events are missing from the seismic
record.
4. The key seismic parameters need to be consistent and reliable. Some seismic source parameters are
estimated using triaxial sensors. If there is only one triaxial sensor in the array, triaxial-based source
parameters will be very dependent on seismic energy radiation patterns and attenuation due to
underground excavations, making the results inconsistent.

For any given seismic event, there will be fluctuations and noise in the seismic source parameters. The issues of
sampling, collection and primary versus secondary failure mechanisms are pertinent. To cope with these
variations, analysis techniques need to consider the trends of the overall populations as well as the parameters of
the larger and more dominant events. The usefulness and applicability of analysis techniques also depends on
factors such as: the quality and type of data available, the experience of the person performing the analysis, and
the fundamental technical basis of the method. A set of attributes and characteristics are suggested to assess the
value of seismic source parameter analysis techniques.

Universal. A seismic source parameter analysis technique should be applicable to more than one specific data
set and for more than one particular mine or mining environment. In addition, the technique needs to be
applicable for real world seismic data, allowing for data noise, and limitations in data sampling of the seismic
record.

Reliable and consistent. A technique should be able to reach similar conclusions for similar sets of data. The
successful application of an analysis technique should not be highly dependent upon specific data filtering,
averaging techniques or other forms of data manipulation. If the results change dramatically as a result of the
calculation process or analysis process, the reliability and consistency of the technique is diminished. Two
independent people should be able to draw the same conclusions using the technique.

Valid. The seismic source parameter analysis technique needs to be able to draw valid conclusions based on
data analysed. It needs to be objective, such that the results of the analysis are unambiguous, in an absolute or
relative context. If the results of the analysis are inconclusive, this should also be clear and incontrovertible.

Auditable. Ideally, a seismic data analysis technique has a means of internal data check. If there are issues with
regard to the quality of the data, or the calculation method, it is helpful to be able to verify that the raw data are
correct and well-behaved.

Practical meaning and interpretation. For a seismic source parameter technique to be useful to understanding
rockmass failure, it is desirable that the analysis technique has a practical meaning or tangible interpretation.
Ideally, the technique can be easily related to general concepts of rockmass stability. A few simple questions to
ask are: “What does this analysis technique mean with regard to changes in a rockmass? Do the results of the
analysis make intuitive sense?”

Marty Hudyma
Assistant Professor, Laurentian University 20
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

2 SEISMIC HAZARD

2.1 Event Magnitude

2.1.1 Description

Magnitude is a measure of the size or intensity of a seismic event. It is usually based on the event amplitude
over a particular frequency range, and generally quantifies radiated energy over a fixed frequency band
(Gibowicz and Kijko, 1994). There are numerous magnitude scales used in understanding mining-induced
seismicity. The Richter Magnitude (Richter, 1935) is the most widely recognised magnitude scale. It was
originally developed based on earthquakes in Southern California. It is widely used to describe the strength of
natural earthquakes. The Geological Survey of Canada reports earthquakes and large mining-induced seismic
events in Eastern Canada using a scale called the Nuttli Magnitude scale (Nuttli, 1973). The Moment Magnitude
Scale (Hanks and Kanamori, 1979) is derived from seismic moment, which is one of the independent seismic
source parameters that can be used to describe a seismic event. Seismic moment is often considered the best
measure of the size of a fault-slip seismic event. With seismic monitoring systems in mines, it is often not
technically possible to calculate a widely accepted magnitude scale, such as Richter magnitude. The term Local
Magnitude is used to describe a magnitude scale calibrated with a local seismic monitoring system.

There are some important characteristics of a magnitude scale include:


 The scale needs to be consistent and reliable, so events of the same size are assessed as being
approximately the same size.
 The scale should include the vast majority of events. If a magnitude scale does not include a
significant proportion of the events recorded, it is less useful for trying to describe a population of
events.
 It should use more than one sensor to assess the event magnitude. If only one sensor is used to
assess magnitude, the assessment will be less accurate, largely due to attenuation or blinding due to
underground excavations, and seismic event energy radiation pattern effects.
 The scale should follow a frequency-magnitude power law relation for the majority of the seismic
data. A reliable power law relation is a crucial characteristic for assessing seismic hazard.

Table 2 contains a qualitative relation between an approximate Richter magnitude of a seismic event and how it
is felt in a mine (Hudyma, 1995). Experience with mine seismicity in several Canadian and Australian mines
was used to develop this relation.

Marty Hudyma
Assistant Professor, Laurentian University 21
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Table 2 Qualitative relation between the magnitude of a seismic event and how it is felt in a mine.

Approximate
Richter Qualitative Description
Magnitude
-3.0  Small bangs or bumps heard nearby. Typically these events are only heard
relatively close to the source of the event.
 This level of seismic noise is normal following development blasts in
stressed ground.
 Events are audible but the vibration is likely too small to be felt.
 Not detectable by most microseismic monitoring systems.
-2.0  Ground shaking felt close to the event.
 Felt as good thumps or rumbles. May be felt remotely from the source of the
event (more than 100 metres away).
 Often detectable by a microseismic monitoring system.
-1.0  Often felt by many workers throughout the mine.
 Should be detectable by a seismic monitoring system.
 Significant ground shaking felt close to the event.
 Similar vibration to a distant underground secondary blast.
0.0  Vibration felt and heard throughout the mine.
 Bump may be felt on surface (hundreds of metres away), but may not be
audible on surface.
 Vibrations felt on surface similar to those generated by a development round.
1.0  Felt and heard very clearly on surface.
 Vibrations felt on surface similar to a large production blast.
 Events may be detected by regional seismological sensors located a few
hundreds of kilometres away.
2.0  Vibration felt on surface is greater than large production blasts.
 The Geological Survey of Canada can usually detect events of this size.
3.0  Event is detected by earthquake monitors throughout the province.
4.0  Largest mining-related seismic events ever recorded in Canada.

2.1.2 Limitations in Magnitude Estimation

Mining people have a general grasp of and are relatively comfortable with the Richter scale, so many mines look
to report a magnitude that can be compared to a Richter magnitude. One of the challenges is that Richter
magnitude is not a typical output from a seismic monitoring system. There have been numerous studies looking
to compare local derived mining-induced seismic event source parameters to the Richter magnitude. Seismic
moment is compared to Richter magnitude for a number of these studies in Figure 13 (Hudyma, 2004).

Observations from Figure 13:


1. There is no data from events smaller than Richter magnitude 1.0, so any relation suggested for
events smaller than Richter magnitude 1.0 would be a pure data extrapolation.
2. The earlier studies (Spottiswoode and McGarr, 1975; Hasegawa 1983) were conducted prior to the
availability of the current commercial seismic monitoring systems. These studies appear to
overestimate the moment magnitude compared to the more recent studies and data.
3. The relations proposed by Kaiser et al. (1996), Talebi et. al. (1997) and Simser et. al. (2003) are
similar, summarise large amounts of data, and fit reasonably well with the majority of the published
Canadian and Western Australian data.

Marty Hudyma
Assistant Professor, Laurentian University 22
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 13 Various studies comparing Richter magnitude to the seismic moment and moment
magnitude source parameters (Hudyma, 2004).

Richter magnitude is a widely used, yet frequently disparaged scale. Richter magnitude is widely used in studies
of peak ground motion and ground support in mines (McGarr, 1984; Kaiser et al., 1996). However, magnitude
and the Richter scale are widely viewed in the seismological community as inferior to other scales and means of
describing a seismic event (Gibowicz and Kijko, 1994). Ultimately, magnitude, or any one other parameter does
not comprehensively describe a seismic event.

Nevertheless, a reliable, well-behaved magnitude scale is the cornerstone of many seismological techniques and
seismic hazard assessment.

Marty Hudyma
Assistant Professor, Laurentian University 23
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

2.2 Frequency-Magnitude Analysis

Frequency-magnitude, or b-value, analysis of seismic events is one of the most widely used techniques for
seismic hazard analysis.

Frequency-magnitude analysis of seismic events was first proposed in earthquake seismology (Gutenberg and
Richter, 1944). It was recognised that the relation between the frequency of occurrence and the magnitude of
seismic events follows a power law relation. For example, for every 1000 events of a magnitude equal to or
greater than -2, we expect 100 events with a magnitude equal to or greater than –1. We also expect 10 events
with a magnitude equal to or greater than 0, and we expect 1 event with a magnitude equal to or greater than +1.

The frequency-magnitude relation can be expressed as:

Log N = a – b m
where,
m = the event magnitude,
N = the number of events equal or greater than magnitude, m,
a = is related to the number of events in the population, and
b = the power law exponent, typically near 1.0 for very large populations of events.

The relation is shown graphically in Figure 14. The magnitude at which the seismic data ceases to follow a
linear, power law relation is usually considered to be the magnitude of completeness of the dataset. The x-axis
intercept of the relation (a/b from Equation 2.6) is widely accepted as a slight overestimate of the largest
magnitude for an event in this population (Gibowicz and Kijko, 1994).
Completeness of the dataset

a/b

Figure 14 A typical Gutenberg-Richter frequency-magnitude relation for a large population of data.


This is a well-behaved dataset, with a linear relation for almost four orders of magnitude,
and a slope that matches a power law.

Marty Hudyma
Assistant Professor, Laurentian University 24
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

When data is plotted on a frequency of occurrence versus magnitude chart, the absolute value of the slope of the
line is the b-value (see Figure 15). The b-value for large populations of seismic events is typically very close to
1.0. While this relation holds true for large subsets of data, when analysing an individual cluster, the b-value
may vary quite dramatically, depending upon the mechanism of the seismicity. Fault-slip related seismicity
frequently has a very low b-value (often less than 0.8). Seismicity as a result of stress change (associated
directly with stress change due to mine blasting) frequently has a b-value in the range of 1.2 to 1.5 (Legge and
Spottiswoode, 1987; Hudyma et al., 1995).

Figure 15 b-value charts from two different seismic sources at a mine in Australia (Hudyma et al.,
2003). The b-value chart on the left is typical of a fault-slip seismic source mechanism.
The b-value chart on the right is typical of a stress change related seismic source
mechanism. The two clusters have approximately the same number of Local Magnitude =
0 events, but the cluster on the left has a relatively high proportion of large events. The
cluster on the right has almost no large events and many small events. The relative seismic
hazard is much higher for the cluster on the left.

b-value really describes the frequency of occurrence of large events versus small events in the dataset. Datasets
with b-values of 0.5 have a much higher proportion of large events than datasets with b-values of 1.5. As such,
b-value is an indicator of relative seismic hazard.

Figure 16 shows the frequency magnitude relations for two groups of events, one group located on a mine fault,
while the other group is near a mine stope (Hudyma, 2008).

Marty Hudyma
Assistant Professor, Laurentian University 25
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 16 Frequency-magnitude relations for two groups of events at an Australian mine (Hudyma,
2008). The fault-slip seismicity has a b-value of 0.5, while stope blasting related seismicity
has a b-value of 0.9.

For a well-behaved (linear) frequency-magnitude relation, the intercept of the x-axis of the frequency-magnitude
relation (which is also equal to the parameters a/b) is sometimes viewed as the maximum expected event size.

Marty Hudyma
Assistant Professor, Laurentian University 26
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

2.3 Magnitude Time History Analysis

2.3.1 Introduction

Magnitude time history analysis (plotting of event magnitude versus time) may be the most insightful technique
for analysing mine seismicity data, giving indications of both seismic hazard and seismic source mechanism. In
a magnitude-time history chart, each of the events is plotted chronologically, and the cumulative number of
events is shown on the secondary y-axis. The slope of the cumulative number of events line is analogous to
seismic event rate, and indicates how the seismic event rate changes over time.

Typical magnitude-time history charts are shown in Figure 17, Figure 18 and Figure 19. There are several useful
pieces of information in this chart.
1. It shows the magnitude of the largest event that has occurred in the seismic monitoring history. The
largest event that has occurred is an indicator of seismic hazard.
2. The number of large events gives an indication of the rockmass damage potential of the seismicity.
3. The timing of events compared to mine blasting can be identified. “Step-wise” behaviour of the
cumulative number of events line is an indication of a strong rockmass response to blasting (Figure 17).
Constant sloped lines are an indication that seismicity is not triggered directly by blasting (as is often
the case with fault-slip seismicity, Figure 19). Changes in slope are indications of a change in the rate
of the rockmass failure process.
4. The timing of significant events compared to mine blasting gives an indication to whether the larger
events are being directly triggered by blasting (as seen in Figure 18). Often for structurally controlled
seismicity, the largest events do not occur at the same time as the majority of events (steps or changes
in slope in the cumulative number of events line).
5. A trend in increasing maximum event size may be an indicator that the failure process is worsening or
the volume of failure is getting larger (Figure 18).

Effectively, the magnitude-time history chart tells a story about the seismicity, the mechanism of the events, and
the seismic hazard.

Marty Hudyma
Assistant Professor, Laurentian University 27
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 17 A magnitude-time history chart of a group of events at Mount Charlotte mine (Hudyma et
al., 2003). The period of nearby stope blasting is represented by the band at the top of the
chart. In this example, there are no large events. This information alone would suggest
that the occurrence of a large seismic event in group of events is unlikely, given that there
has not been an event greater than magnitude 0. It can also be seen that the events in this
cluster occurred over a fairly short time period (less than 9 months). The events located in
the back of a large open stope. Prior to stope blasting there were no events in the cluster,
and when stope blasting stopped the events stopped. These are typical characteristics of a
source mechanism that is dominated by induced stress change due to mining. It is
reasonable to conclude that the relative seismic hazard in this cluster of events is low.

Marty Hudyma
Assistant Professor, Laurentian University 28
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 18 A magnitude-time history chart of another group of events at Mount Charlotte mine
(Hudyma et al., 2003). The period of nearby stope blasting is represented by the band at
the top of the chart. The largest event recorded in group is local magnitude 1.5, and there
have been a few events that could be described as potentially damaging. These large
events are spread over a period of more than a year and records show that the events are
mainly unrelated to specific stope blasting. These are typical characteristics of a source
mechanism that is related to a fault. A main fault at Mt. Charlotte mine crosses through
this group of events. Due to the number of larger events, the seismic hazard associated
with this seismic source in Figure 18 is significantly higher than the one in Figure 17.

Marty Hudyma
Assistant Professor, Laurentian University 29
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 19 Magnitude-time history plot of the seismicity on a mine fault (Hudyma et al., 2003). There
was blasting almost weekly over the 12-month period, but the seismic response from the
fault is constant, suggesting that the rate of seismicity is unaffected by mine blasting.

2.3.2 Other Considerations

The cumulative number of events is a means of investigating changes in seismic event rate. This parameter is
only meaningful if the seismic system sensitivity remains relatively constant. For example, Figure 20 is a
Magnitude-Time History chart for all of the seismic events recorded in a mine between October 2003 and June
2006. The rate of seismic events changes significantly over the time period. There are significant increases in
event rate in April and August 2004 and a significant decrease in event rate in July 2005 (denoted by red dashed
lines in Figure 20). These changes could be interpreted as mining-induced changes in the rate of seismic events
due to stope extraction in the mine.

Marty Hudyma
Assistant Professor, Laurentian University 30
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 20 Magnitude-Time History chart for all of the events recorded in a mine between October
2003 and June 2006.

However, the first event rate increase (April 2004) was due to a seismic monitoring increase in sensitivity that
occurred with the addition of more seismic sensors. The July 2005 change in event rate is a consequence of a
change in standards for processing seismic data (the mine no longer had a full time person dedicated to
processing seismic data), which resulted in smaller seismic events not being located. Figure 21 uses frequency-
magnitude charts to show how the seismic system sensitivity changed from –0.5 (from October 2003 to April
2004), to –1.1 (from April 2004 to July 2005), to –0.8 (from July 2005 to June 2006). The changes in seismic
system sensitivity can also be identified from the minimum event magnitude in Figure 20. The minimum event
magnitude from October 2003 to April 2004 is approximately magnitude –1.5. There are only a couple of events
smaller than magnitude –1.5. Between April 2004 and July 2005, the minimum event magnitude is
approximately –2.5. From July 2005 to July 2006, the minimum event magnitude is approximately –2.0, with
only a few events smaller than magnitude –2.0 recorded. The Magnitude-Time History chart shows that the
system sensitivity is changing, and that these changes are affecting the rate at which seismic events are being
recorded.

The influence of seismic system sensitivity can be removed from Magnitude-Time History charts by plotting the
events only greater than the seismic system sensitivity. Figure 22 is a Magnitude-Time History chart for the
events with a magnitude of –0.5 or greater. The chart shows that the April 2004 and July 2005 event rate
changes were artificial, as there was no change in event rate for the larger events. The July 2004 event rate
change is a real increase in seismic event rate (due to the onset of sill pillar mining at the mine).

Marty Hudyma
Assistant Professor, Laurentian University 31
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

b = -1.1 b = -1.2 b = -1.2

Sensitivity
Sensitivity
Sensitivity

Figure 21 Seismic system sensitivity for October 2003 to April 2004 (left), April 2004 to July 2005
(middle), and July 2005 to June 2006 (right). The seismic system sensitivity changes from
–0.6 (left), to –1.1 (middle), and –0.8 (right).

Figure 22 Magnitude-Time History for events of magnitude –0.5 and larger.

Finally, a broad population of events is one that encompasses a significant portion of a mine, incorporating
seismicity from many different rockmass failure processes. When a broad population of seismic events is
plotted in a Magnitude-Time History chart, the technique will give less detailed information about seismic event
failure processes. With numerous rockmass failure processes and seismic source mechanisms in a broad
population, limited insight can be gained into the seismic source mechanism using Magnitude-Time History
analysis. In addition, it may be more difficult to identify cause-effect relations between seismicity and mine
blasting. Magnitude-Time History analysis provides more insight into mining-induced seismicity when spatially
constrained populations of data are analysed.

Marty Hudyma
Assistant Professor, Laurentian University 32
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

2.4 Instability Analysis

Instability analysis, also called Energy Index Cumulative Apparent Volume analysis, is a model dependent
seismic analysis technique that is occasionally used to proactively identify major rock mass instability.

Instability analysis involves comparing two seismic source parameters, energy index and cumulative apparent
volume. Energy index (EI) is a parameter related to concentration and accumulation of stress in a rock mass.
For populations of seismic events, there is a logarithmic relation between seismic energy released and seismic
moment. EI the relative amount of energy released for a seismic event compared to the amount of energy that
would be expected for an event of that size (Figure 23). An EI of more than one signifies that more energy has
been released that expected, while an EI of less than one signifies that less energy has been released that
expected. Past studies (van Aswegen and Butler, 1993) have found that when an area with a structure (pillar,
fault or dyke) is accumulating stress the EI for related seismic events is greater than one. When that structure
starts to fail, the EI for the related seismic events drops below one. Typically EI is calculated as an average of a
previous number of events to smooth out large variations.

Figure 23 Calculation of Energy Index (after Mendecki et al., 1999).

Cumulative apparent volume (CAV) is a measure of the rock mass deformation occurring at the time of a
seismic event. Past studies have related CAV to measured rock mass deformation (Minney et al., 1997).

EI and CAV are plotted together on a time history chart looking for changes in the general stability of a
rockmass. Instability starts to occur when there is a drop in EI, signifying that stress is being shed away from a
structure. At the same time, deformation (quantified by CAV) starts to increase significantly. In South African
mines, a drop in EI and an increase in CAV have sometimes been found to be a precursor to very large seismic
events.

Figure 24 shows the seismic behaviour of a cluster of events in a Western Australian mine (Australian Centre for
Geomechanics, 2005). There are two clear examples of instability developing in Figure 24, in mid September
and early December. In both cases, there are significant fluctuations in energy index (stress accumulation),
followed by sudden drops in energy index. Co-incidental with the energy index drops, there are large jumps in
cumulative apparent volume, corresponding to seismic events larger than Richter +1. So, in both cases
Instability Analysis identified unstable failure.

Marty Hudyma
Assistant Professor, Laurentian University 33
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

The instability concept is based on a rock mass failure model, of stress accumulation, followed by large scale
rockmass deformation. This failure model has not been widely validated outside of South African “reef-style”
gold mines, and instability analysis is not widely used outside South Africa to predict large seismic events.
However, the components of instability analysis are sometimes used to investigate whether stress is
accumulating in a structure, or whether significant deformation is occurring in a region, or along a structure.

M=+1.1 M=+1.2

Figure 24 Two examples of classic instability in a cluster of events can be identified using EI/CAV
analysis (ACG, 2005). In both cases, there are large fluctuations in energy index,
indicating stress accumulation followed by drops in energy index shortly before Richter +1
events.

Marty Hudyma
Assistant Professor, Laurentian University 34
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

2.5 Apparent Stress Time History

Past work has shown that seismic events associated with increasing stress typically have a higher seismic energy
than expected (van Aswegen and Butler, 1993; Simser et al., 2003). Instability analysis utilises this concept to
identify regions of increasing and decreasing stress in an effort to identify unstable rockmass conditions that are
potentially prone to large damaging seismic events (Mendecki et al., 1999).

Apparent stress is a model independent measure of the stress change at a seismic source (Mendecki and van
Aswegen, 2001). Originally proposed by Wyss and Brune (1968), apparent stress is defined as:

σa = μ E / Mo
where
σa = apparent stress
μ = the shear modulus of rigidity of the source material,
E = seismic energy, and
Mo = seismic moment.

Figure 25 shows a energy-moment relation for 19280 events recorded over a period of 3 years in an Australian
mine. Figure 26 (left) shows the events that have a high Energy Index (in this case high Energy Index is the
actual energy recorded is 5 times greater than the expected energy). Figure 26 (right) shows the events that have
a high Apparent Stress (in this case high Apparent Stress is arbitrarily chosen to be greater than 50,000 Pascals).
Some observations about the two populations:
 Of the 19280 events, 2428 have high apparent stress and high Energy Index. So to a degree the two
methods of identifying higher energy events are similar.
 71% of the high Energy Index events have a local magnitude less than -1, while only 4% of the high
Energy Index events have a local magnitude greater than 0. So the high Energy Index population is
dominated by smaller events, as would be expected due to frequency-magnitude relation.
 40% of the high Apparent Stress events have a local magnitude less than -1, while 15% of the high
Apparent Stress events have a local magnitude greater than 0. The high Apparent Stress population
is more related to larger events than the high Energy Index events.
 For the significant events (local magnitude ≥ 0), only 208 of the 687 (30%) had a high Energy
Index, while 417 of the 687 (61%) had a high Apparent Stress.

Effectively, trends in Energy Index are more likely to be related to the occurrence of smaller events, while trends
in Apparent Stress incorporates a much higher proportion of the larger events. A seismic source parameter
technique that utilises Apparent Stress is more likely to be related to elevated stress levels and the occurrence of
larger events (elevated seismic hazard).

Apparent Stress Time History

Apparent Stress Time History (ASTH) is a simple parameter. It is the daily number of events with an Apparent
Stress equal or greater than a threshold, in a trailing (preceding) time period. A threshold value of 10,000 to
30,000 Pascals is often a good threshold (usually identifying the top 10 to 20 percent of high Apparent Stress
events). A trailing time period of 3 to 7 days generally produces a good trend, being sensitive to variations in the
number of high Apparent stress events while not being too erratic.

Marty Hudyma
Assistant Professor, Laurentian University 35
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 25 Energy – Moment relation for 19280 events recorded over a period of 3 years in an
Australian mine.

Figure 26 The left diagram shows the events with a high Energy Index (EI ≥ 5). The right diagram
shows the events with a high Apparent Stress (Apparent Stress ≥ 50,000 Pascals).

Examples of the application of Apparent Stress Time History

Zone 7 at Laronde is a small lens to the west and in the footwall of the main orebody. The lens is being mined
with a pillarless retreat open stoping method. ASTH for the Zone 7 is shown in Figure 27. The Apparent Stress
Frequency is averaged over the previous 7 days. The apparent stress threshold is 10,000 pascals. The timing of
19 final stope blasts are also shown in Figure 27.

Marty Hudyma
Assistant Professor, Laurentian University 36
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

1 2 3 4 5 6 7 8 9 10 11 1213 14/15 16 17 18/19

Figure 27 Apparent stress time history for the seismic events around Zone 7 at Laronde.
Observations from Figure 27:
 The background Apparent Stress Frequency is between 0 and 0.5 in Figure 27. Each of the instances
that the Apparent Stress Frequency exceeded 2.0 was due to a stope blast. This occurred for 13 of
the 19 stope blasts. There is a strong relation between stress change due to stope blasting and
Apparent Stress Frequency.
 Within about 1 week of the stope blast (the averaging period used in this case), the Apparent Stress
Frequency drops below 1.0. This suggests that the temporal increase in Apparent Stress drops
quickly after the blast as stress is redistributed.
 For each of the 19 stope blasts, shows there is a general relation the level of Apparent Stress
Frequency and the number of significant events (local magnitude ≥ -1) recorded within a few days
of blasting. This is an expected relation as larger events have a high apparent stress. In this case, it
does show that the level of Apparent Stress Frequency is somewhat related to temporal increases in
seismic hazard.
 This generalisation is further supported by the fact that of the 7 of the 10 largest events, recorded in
Zone 7, occurred when the Apparent Stress Frequency was 3.0 or higher.

In the following 4 examples in another Canadian mine, ASTH is calculated using a threshold of 10,000 Pascals
and using a 7-day trailing average. Time periods in which the daily number of high Apparent Stress events
exceeds 5 is highlighted in red.

Conclusion

The conclusion from these examples is that ASTH is strongly related to stress change due to mine blasting. In
some cases, particularly in high stress mining environments, high levels of Apparent Stress Frequency may be
good indicators of elevated seismic hazard. The optimum Apparent Stress threshold and trailing time period are
site dependent parameters.

Marty Hudyma
Assistant Professor, Laurentian University 37
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 28 In this group of events, all 8 of the large events (local magnitude ≥ 0) occur during or very
close to period in which the Apparent Stress Frequency exceeds 5 events per day.

Figure 29 There are seven periods of mining in the 10 months of this time history (spikes in Apparent
Stress Frequency). There were only two brief periods in which Apparent Stress Frequency
exceeded five events per day. The duration of in which the Apparent Stress Frequency
exceeded 5 events per day was 3 weeks out of a total of 46 weeks. The only large seismic
event occurred during one of the periods of high Apparent Stress frequency.

Marty Hudyma
Assistant Professor, Laurentian University 38
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 30 In this example, there are numerous brief increases in Apparent Stress Frequency. All five
of the large events (local magnitude ≥ 0) occurred during or close to periods of Apparent
Stress Frequency of 5 events per day or greater.

Figure 31 Three of the four large events (local magnitude ≥ 0) occur during or very close to periods
of an Apparent Stress Frequency of 5 events per day. The Apparent Stress Frequency was
high for 3 of the 12 months. This is a particularly interesting example because the seismic
event frequency was relatively high throughout the year.

Marty Hudyma
Assistant Professor, Laurentian University 39
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

2.6 Daily Histogram of Event Frequency

A histogram of the number of event recorded per day is occasionally used as an indicator of the relative level of
seismic hazard in a mine. This is somewhat true when the data is compared relatively at the same mine.
However, more correctly, the number of events per day is more a function of the number of seismic sensors
operating in the mine. If more sensors are installed, more events will be recorded. Nevertheless, there is some
value in utilising histograms of event frequency.

The number of events recorded per day can be used to identify what is a normal level of seismicity versus an
abnormal level of seismicity. Figure 32 is a daily histogram of events at a Canadian mine over a period of two
years. The chart also has a 30-day trailing average of the number of events per day. For this mine, there are
periods in which more than 500 events per day are recorded, while on other days the number of events is less
than 10 per day. The 30-day trailing average suggests the number of events per day usually varies between 20
and 80, giving an average of about 50 events per day. There are “build-ups” of seismicity starting in November
2004 and May 2005 related to stope mining in a highly stressed part of the mine. Once the stoping is complete
in these areas, the number of events quickly returns back to a normal rate.

This chart also quickly shows the completeness of the seismic record. There are periods of several days in
November 2005 and December 2005, when no events were recorded. These are period in which the seismic
system was not operational. There are also brief gaps in the seismic record in July 2004, December 2004, and
February 2006. Overall, the seismic record at this mine is relatively complete.

Figure 32 Histogram of the number of events per day in a Canadian mine.

Marty Hudyma
Assistant Professor, Laurentian University 40
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

3 SEISMIC SOURCE MECHANISM

3.1 S:P Energy Ratio Analysis

3.1.1 Background

For every seismic event, there is a p-wave (also called the compressional wave), and an s-wave (also called the
shear wave). The p-wave velocity in hard rock mines is typically about 6000 metres per second, while the s -
waves in hard rock is approximately 3500 metres per second. Figure 33 shows a typical seismic waveform and
the P-wave arrival and the S-wave arrival (redraw from Hudyma et al., 2003).

Figure 33 Typical triaxial waveform, showing the P-wave and S-wave arrival times (redraw from
Hudyma et al., 2003).

Seismic energy is proportional to the integral of the square of the vectoral sum of the velocity waveform and can
be calculated for the p-wave and s-wave. For fault-slip type events, there is considerably more energy in the s-
wave than in the p-wave (Boatwright and Fletcher, 1984), with the ratio of the s-wave energy to the p-wave
energy frequently greater than 10. Urbancic et al. (1992) note that for non-shear seismic source mechanisms,
there is a deficiency in s-wave energy, or relatively more P-wave energy than for shearing events. For non-
shearing event mechanisms such as strain-bursting, tensile failure and volumetric rock mass fracturing, the ratio
of s-wave energy to p-wave energy is frequently in the range of 1 to 3.

Typically, s-wave to p-wave energy is plotted as scatter plots, as shown in Figure 34 (Hudyma et al., 2003).
These diagrams can be difficult to interpret. The same data is presented in Figure 35 as cumulative distribution
plots. More than 60% of the events in the group on the left have an S:P energy ratio of greater than 10, while
only slightly more than 20% of the events in the group on the right have an S:P energy ratio greater than 10. The
group on the left in Figure 34 has a much higher proportion of high S:P energy ratio (shear events) than the
group on the right. Cumulative distribution plots make it easier to quantify the differences in populations of S:P
energy ratio data.

Marty Hudyma
Assistant Professor, Laurentian University 41
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 34 Scatter plots of s-wave and p-wave energy for two different groups of events (after
Hudyma et al., 2003) . The left group (a) has more shear related events (S:P>10) than the
right group (b), but the busy nature of the scatter plots makes them difficult to interpret.

Figure 35 Cumulative distribution of the two scatter plots shown in Figure A8 (Hudyma et al., 2003).
The left group in Figure A8 is shown with a solid line, while the right group in Figure A8 is
shown with a dashed line. The left group has a much higher proportion of high S:P energy
ratio (shear) events.

Figure 36 and Figure 37show the distribution of s-wave to p-wave energy for two clusters of seismic events for
Mount Charlotte mine in Australia (Hudyma et al., 2003). In Figure 36, the fault slip cluster of events has a
much higher ratio of s-wave to p-wave energy than the cluster of events whose mechanism is primarily non-
shear (Figure 37).

Marty Hudyma
Assistant Professor, Laurentian University 42
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 36 The distribution of S:P energy for a cluster of seismic events (dashed line) compared to all
of the events in at Mt. Charlotte mine (Hudyma et al., 2003). The median (50%) S:P
energy ratio for the mine is about 6.5, which suggests that the majority of events in the
mine have a high S-wave energy. The median S:P energy ratio for Cluster #18 is
approximately 14. This suggests that events in Cluster #18 are strongly shear related.
Cluster #18 is directly on the Neptune fault at Mt. Charlotte (Hudyma et al., 2003).

Figure 37 The distribution of S:P energy for a different cluster of seismic events (dashed line)
compared to all of the events in at Mt. Charlotte mine (Hudyma et al., 2003). The median
(50%) S:P energy ratio for the mine is about 6.5, while the median S:P energy ratio for
Cluster #10 is approximately 2. This suggests that events in this cluster are strongly non-
shear related (Hudyma et al., 2003).

Marty Hudyma
Assistant Professor, Laurentian University 43
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

3.1.2 Additional Considerations

Scale dependence is a fundamental consideration when analysing seismic source mechanism. It is important to
differentiate whether the small events have the same seismic source mechanisms as the larger events. The data
in the scatterplot in Figure 34(b), is shown in Figure 38. Isolines of S:P energy are shown for 0.1, 1.0, 10, and
100. There is a clear change in the data. Above 20 joules of P-wave energy, there is only one seismic event with
an S-wave to P-wave energy ratio of less than three. For P-wave energy less than 20 joules, there is a much
wider range of S-wave to P-wave energy, varying from 0.3 to 30.

Figure 38 S-wave versus P-wave energy for 490 events in pillar at Perseverance mine (Australia).
Lines of equal S:P energy ratio are shown for S:P energy of 0.1, 1.0, 10, and 100.
The same data is shown divided into magnitude ranges in a cumulative frequency distribution in Figure 39.
Observations from Figure 39:
 The smallest events (–3≤ML<–2) have a much lower S:P energy ratio than the rest of the events.
The median S:P energy is about four, and only 11% of the events have an S:P energy ratio
greater than 10 (typical shear events).
 The two ranges of largest events (–1≤ML<0, 0≤ML<+1) have very similar S:P energy
distributions. The median S:P energy ratio is about 7.5, and about 31% of events have an S:P
energy ratio greater than 10.
 The (–2≤ML<–1) magnitude range is between the two extremes, with a median S:P energy ratio
of about 6.5 and about 20% of event having an S:P energy ratio greater than 10.
 It can be concluded that there is a significant magnitude dependence of S:P energy ratio in the
population of events, with the events smaller than –2 being of a potentially different or
secondary seismic source mechanism.

Marty Hudyma
Assistant Professor, Laurentian University 44
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 39 S:P frequency distributions for different magnitude ranges for the 490 events in Figure 38.
Events directly induced by mine blasting may be the source of the scale dependence. Smaller events (ML < –1)
are potentially caused by stress change due to blasting. The larger events are less likely to be blast related,
having a shearing seismic source mechanism.

Scale dependence of S:P energy ratio was investigated for seismicity along the the Big Bell Fault at Big Bell
(Hudyma, 2008). The faults is more than 400 metres from active stoping, so there is minimal direct influence of
stope blasting on the occurrence of seismic events.

Figure 40 S:P energy ratio for the Big Bell Fault at Big Bell mine in Western Australia.

Marty Hudyma
Assistant Professor, Laurentian University 45
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

There is some variation in the median S:P energy ratio and the percentage of events with an S:P energy ratio
greater than 10. However, almost all magnitude ranges of events have high S:P energy ratios, suggesting that
shearing is the dominant seismic source mechanism. The percentage of events with low S:P energy ratios varies
between 5% and 20%.

It cannot be generalised that scale dependence of S:P energy ratio always has a high S:P energy ratio for larger
events. Figure 41 shows the distribution of S:P energy ratio for seismic events in a highly stressed mine
abutment. There is a clear scale dependence in the S:P energy ratio of this population of events. In this
example, the median S:P energy ratio decreases with increasing event magnitude. For smaller events (magnitude
less than –1.0), the seismic source mechanism is more shear related with 40% to 50% of events having an S:P
energy ratio greater than 10. For events larger than magnitude –1.0, the median S:P energy ratio is
approximately three, with less than 15% of the events having a high S:P energy ratio. The main seismic source
mechanism for the larger events is inferred to be stress related volumetric fracturing.

Figure 41 S-wave to P-wave energy ratio for a highly stressed abutment.

Marty Hudyma
Assistant Professor, Laurentian University 46
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

3.2 Diurnal Analysis

Diurnal (or time of day) analyses were first proposed by Cook (1976) to compare differences in seismicity
between mines. Seismicity directly associated with stress change due to mine blasting occurs directly after mine
blasting. Seismicity associated with geological structures often has little temporal relation to mine blasting. For
example, in the population of events in Figure 42, more seismic events occur at blast time at 05:00 and 17:00.
These events occurred in a highly stress abutment of a mine stope. In Figure 43, seismic events do not
preferentially occur at a specific time of day. These events occurred along a mine fault, about 150 metres from
the nearest mine stope.

Consequently, diurnal analysis can be an indicator of seismic source mechanism.

Figure 42 Seismic events are occurring more frequently directly following mine blasting at 05:00 and
17:00. These events are occurring in a highly stressed mine abutment.

Figure 43 Seismic events occur are not occurring at a specific time of day for this group of events on
a mine fault.
As an additional example, several thousand seismic events were recorded associated with the development of a
decline in a mine. Two populations of seismic events were identified, one population associated with decline
mining in competent basalt and one population associated with decline mining through stiff porphyry dykes.
Diurnal charts for the two populations are shown in Figure 44 and Figure 45, respectively. For the basalt decline
population, the average number of events per hour is 187, with a standard deviation of 112. Lines are shown on

Marty Hudyma
Assistant Professor, Laurentian University 47
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 45 for the average number of events as well as the average, plus and minus one and two standard
deviations. There are no hours in which the number of events is less than one standard deviation below the
average, but there are four hours in which the number of events per hour exceeds one standard deviation above
the average. For 04:00, the number of events is more than two standard deviations above the average.
Development mining blasts are typically fired at approximately 04:00 and at 17:00 each day.

The results are fairly similar for the porphyry dyke population. There is a large increase in seismicity at blast
times at 04:00 and 17:00. However, there is a greater peak of seismic activity at 04:00 for the basalt decline
events compared to the porphyry decline events. This would suggest that events in the basalt rockmass unit are
more likely to occur at blast time than in the porphyry dyke.

Figure 44 Diurnal analysis of events associated with development mining in the basalt rockmass unit.

Figure 45 Diurnal analysis of events associated with development mining in porphyry dykes.

Marty Hudyma
Assistant Professor, Laurentian University 48
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

3.3 Phasor Analysis

Timing of seismic events can also be investigated using phasor diagrams. Phasor diagrams are based on the
Schuster test (Rydelek and Sacks, 1989). In the Schuster test, the time (phase) of each event is plotted as a unit
length, in an X-Y chart. The phase of all of the events is combined to create a phasor sum. If the times of events
are truly random, the phasor sum is also random. If the time of events is not random, the phasor sum exhibits a
directional walkout. Rydelek and Hass (1994) proposed the use of phasor diagrams to determine if a significant
number of blasts were contained within earthquake seismic catalogues. Figure 46(a) is a phasor diagram for all
of the events recorded at Big Bell.

An extension of this technique is to identify datasets that contain a large proportion of events directly induced by
mine blasts. The technique needs to be modified if more than one central blast time is used at the mine. For
mines that have two central blasting periods, the phasor diagram needs to be modified from a 24 hour chart to a
12 hour chart, to avoid the two blast times having a cancelling effect. For example, phasor diagrams for the
entire seismic catalogue (more than 24,000 events) are shown for 24 hours and 12 hours (Figure 46).

Figure 46(a) shows a phasor walkout direction towards 06:00. Over time, the phasor path starts to move towards
18:00. The phasors of the 06:00 blast-related events and 18:00 blast-related events essentially cancel each other.
When the phasor diagram is changed to 12 hours (to prevent 06:00 and 18:00 cancelling), the phasor shows a
very strong walkout at 06:00/18:00, which is more representative of the high seismic event rate directly after
blasting. Rydelek and Sacks (1989) note that if a seismic record is random (for a confidence of 95%), the
random walkout length, R, should not exceed 1.73 N , where N is the total number of events. The random
walkout length, R, is shown by a circle in the phasor diagrams in Figure 46.

03:00

(a) (b)
15:00

06:00/18:00
00:00/12:00

09:00
21:00

Figure 46 Phasor diagrams for all of the events (24,077 events) using a 24 hour phasor chart (a) and
a 12 hour phasor chart (b).
It has been shown that diurnal charts and phasor analysis can be used to infer whether the seismic events in a
population are being directly induced by mine blasting or whether the timing of events is independent of mine
blasting. Events that occur independent of mine blasting are more likely to be related to geological features.

There are important scale dependence issues. Phasor analysis of the events in the stiff porphyry dykes at in a
Western Australian mine shows there is a clear relation between mine blasting and the occurrence of the smaller
seismic events smaller than magnitude –0.5 (Figure 47). However, for events greater than magnitude –0.5, there
is no longer a statistically significant occurrence of seismicity at blast time.

Marty Hudyma
Assistant Professor, Laurentian University 49
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

(a) (b)
All events M -1
N=1383 N=269
R=64.3 R=28.4

(c) (d)
M -0.5 M 0.0
N=67 N=14
R=14.2 R=6.5

Figure 47 Phasor analysis of the events in a stiff porphyry dyke in a Western Australian mine.

Larger events in a population sometimes do show a statistically significant frequency of occurrence at blast time.
Phasor charts for the highly stressed mine abutment at are shown in Figure 48. There is a significant walkout for
all events as well as events larger than magnitude 0.0.

(a) (b)
All events M 0
N=4770 N=114
R=119.5 R=18.5

Figure 48 Phasor analysis of the events in the highly stressed mine abutment.

Marty Hudyma
Assistant Professor, Laurentian University 50
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

4 DATA INTEGRITY

The data analysis techniques described above have a number of means of data verification. If the data quality is
low, or the seismic record is short, or the number of seismic events involved in the analysis is limited, the
validity of the analysis may be affected. The following section shows how histograms, Magnitude-Time History
analyses, energy-moment relations, and frequency-magnitude relations can be used to identify seismic data that
may be tainted by outliers or corrupt data.

4.1 Daily Histogram

A histogram of the number of events per day can be useful in checking that the seismic data record is relatively
complete. Figure 49 is a histogram of the number of events per day over a three year period. There are several
days in which there are no events recorded (arrows in Figure 49). There appears to be a few days of data
missing in August 2004, February 2006 and July 2006. There are likely several days of data missing in
November 2005. The event rate in March 2005 is very low compared to the rest of the seismic record, there may
be some partial data loss in this period, or it may have been a very quiet time in the mine. Overall, this is a
relatively complete seismic record.

In Figure 50, the number of events per day is shown over a period of about 1.5 years. There are numerous
periods in which there is no data (red arrows), and there are a number of periods in which the number of events
recorded is very low (green arrows). Periods in which the number of events is very low may be due to a
temporary decrease in seismicity in the mine. However, it is more likely this represents time periods when a few
or several seismic sensors were not operational. In this example, the seismic record is likely inadequate,
potentially affecting the seismic source parameter analyses and seismic data interpretations.

Figure 49 Histogram of the number of events recorded per day. The seismic record is relatively
complete, with several days of data loss in November 2005, and possibly a few days of
missing data in August 2004, February 2006, and July 2006.

Marty Hudyma
Assistant Professor, Laurentian University 51
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 50 The histogram of the number of events per day shows many period of no seismic activity
(denoted by red arrows), and several periods in which the number of event recorded
appears to be too low compared to the overall seismic record (denoted by green arrows).
In this example, the seismic record appears to be quite incomplete and likely inadequate
for detailed back-analysis.

4.2 Magnitude-Time History

A Magnitude-Time History chart can be used to check a number of aspects of the seismic record. The
cumulative number of events in Figure 51 is flat in April, May, October, and November 2004 and April 2005.
There are periods of data missing in these months.

A three year seismic record in Figure 52 has significant changes in event magnitude. Red dotted lines have been
put on Figure 52 to show the approximate minimum and maximum event magnitude being recorded. From
January 2005 to August 2006, the smallest events recorded were approximately magnitude –3.8 and the largest
events being recorded were approximately magnitude +0.8. From August 2006 to April 2007, the minimum
event size dropped to –5.0 and largest event size was approximately –0.2. This shift in the magnitude scale
appears to represent a change in how event magnitude is being calculated. There is likely an analysis software
change at this time. In April 2007, there may have also been a change in the magnitude scale, as the minimum
event size and the maximum event size both appear to have increased slightly. The change in magnitude scale in
April 2007 is a fundamental change in the seismic record. The magnitude of events before and after April 2007
cannot be compared. Importantly, it is likely that other primary seismic source parameters, such as seismic
energy and seismic moment, may have also been affected by this change in the seismic system.

Marty Hudyma
Assistant Professor, Laurentian University 52
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 51 There is missing data in the seismic record in April, May, October, and November 2004
and April 2005.

Figure 52 There are changes in the magnitude scale August 2006 and possibly again in April 2007.
There is a seven year seismic record shown in Figure 53. The minimum event size during this time period
changes subtly several times (shown with red dashed lines in Figure 53); however the maximum event size
appears to remain relatively constant. Likely these changes were caused by temporary inoperation of some of
the seismic sensors in the array. Alternatively, the variations may represent a different level of diligence in the
seismic system operators in processing the seismic data.

Magnitude-Time History analysis can also be useful to identify events that potentially do not belong with a
population. Figure 54 shows a population of seismic events that were recorded over a period of several years.
The largest seismic event was recorded as magnitude +1.5. The event occurred at a time of relatively low event
rate, suggesting that there was no external influence (mine blast or nearby rockmass failure) to cause the seismic

Marty Hudyma
Assistant Professor, Laurentian University 53
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

event. There were only five events greater than magnitude 0.0 recorded in this population. The largest seismic
event is much larger than the rest of the population and suggests that the event:
 Is poorly captured and may have unreliable seismic source parameters.
 May be poorly located and should not be associated with this population of events.
 Or, is actually a mine blast that was inadvertently included in the seismic event record.
When the magnitude of one or a few events appear to be unduly high, compared to the rest of the population of
events, the events should be checked to ensure the locations and parameters are correct, and to ensure that the
event is not a mine blast.

Figure 53 Over a seven year seismic record, there are subtle changes in the minimum magnitude of
seismic system. These changes likely reflect periods in which some of the sensors were not
operational and the seismic system was less sensitive. The variations may also reflect
changes in the diligence in the seismic system operators in processing seismic data.

Figure 54 Magnitude-Time History analysis shows a single event that does not fit the character of the
other seismicity in the population of events.

Marty Hudyma
Assistant Professor, Laurentian University 54
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

4.3 Energy-Moment Relation

There is a logarithmic relation between the energy radiated from a seismic event and the seismic moment of an
event. A typical “energy-moment” relation is shown in Figure 55. The slope of the relation is usually in the
order of 1.2 to 1.6. The intercept of one joule of seismic energy is usually in the order of 10,000,000 Nm
(seismic moment). In this population, seismic energy varies up to plus or minus an order of magnitude from the
best-fit relation (blue line in Figure 55). This is a well-behaved population of events. The relation can be
expressed as:

log(E)  c  d  log(Mo)

where,
E = seismic energy (joules)
Mo = seismic moment (Nm)
c = is a constant related to the stress acting on the rockmass system (Mendecki and van Aswegen, 2001)
d = is a constant related to the stiffness of the rockmass system (Mendecki and van Aswegen, 2001)
Energy (joules)

Moment (Nm)
Figure 55 A typical energy-moment relation for a seismic data set.

The energy-moment relation can be a useful tool for identifying problems or trends with a seismic data
population. Figure 56 is the energy-moment relation for a population of seismic events recorded at a mine.
There are two clear distinct populations of data. While different rockmasses may have slightly different
relations, it is unlikely that such a marked change is a natural phenomenon. Likely there has been a fundamental
change in the settings or software of the seismic monitoring system, similar to that shown in the Magnitude-
Time History chart in Figure 52.

Marty Hudyma
Assistant Professor, Laurentian University 55
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Energy (joules)

Moment (Nm)

Figure 56 There are two clear populations of seismic data in this dataset.

Figure 57 shows a population of seismic events, with the events plotted by S-wave to P-wave energy ratio. The
smaller events (lower seismic energy and seismic moment) have a much higher S:P energy ratio compared with
the larger events. There is a scale dependence of S:P energy ratio in this data set. It should also be noted that
this population has been filtered based on a combination of energy and moment (dotted line in Figure 57),
creating a truncated dataset.
Energy (joules)

Moment (Nm)

Figure 57 The seismic events in this population are plotted by S:P energy ratio.

Marty Hudyma
Assistant Professor, Laurentian University 56
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 58 shows a widely dispersed energy-moment relation. A regression best-fit is shown with the blue line,
but there could potentially be two separate populations in the data, shown with dashed black lines. Also, for any
given seismic moment, the events with a higher seismic energy have a higher S:P energy ratio. This may
represent a natural characteristic of the seismic data, but may also be an indicator of two separate populations of
events.
Energy (joules)

Moment (Nm)

Figure 58 There may be two populations of events shown in this dataset. There is also a clear
gradation of S:P energy ratio in this dataset.

4.4 Gutenberg-Richter Frequency-Magnitude Relation

The Gutenberg-Richter frequency-magnitude relation can be used as a tool to check data integrity. A dataset
would be expected to have a linear relation over a relatively large range (usually two or more orders of
magnitude), and the slope of the data should be near –1.0, following a power law relation.

The data in Figure 59 does not follow an expected power law. The slope of the frequency-magnitude relation is
–1.6 rather than the expected –1.0. For this data, it is suspected that the magnitude scale is poorly defined or
corrupted by some aspect of the seismic monitoring hardware or software.

Marty Hudyma
Assistant Professor, Laurentian University 57
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Figure 59 The frequency-magnitude relation is not following an expected power law. The magnitude
scale being used for the seismic data is potentially unreliable.

In Figure 60 the frequency-magnitude is not linear over a significant range. A linear relation can be drawn
between magnitude +0.5 and +1.0, however, it is linear for less than an order of magnitude, and the relation does
not follow a power law. In this case, the magnitude scale is dubious and untrustworthy.

Figure 60 The frequency-magnitude relation is not linear. The best linear fit to the data is between
magnitude +0.5 and +1.0. Typically, the data should follow a linear relation for two or
three orders of magnitude, and have a near power law relation (the slope should be about
–1.0, not –1.8).

In Figure 61, the data follows a good frequency-magnitude relation with a linear power law for two orders of
magnitude. However, the largest event in the dataset is magnitude +1.5, is significantly greater than the
expected largest magnitude. This may occur for a number of reasons:
 The monitoring period may be short and not sufficient to be adequately representative of the sources of
seismicity represented by the data.

Marty Hudyma
Assistant Professor, Laurentian University 58
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

 There may be a tectonic component or influence in the data set that is not adequately represented by the
data recorded. This can occur when a major fault exists within the monitoring area and ruptures during
nearby mining.
 The dataset is filtered in time and or space and the very large event is part of a rockmass failure process
excluded by the filtering, i.e. the event is poorly located and does not belong in this population. This is
a relatively common problem as very large events are frequently poorly located by a seismic monitoring
system.
 The parameters of the large seismic event may be erroneous, caused by the poor collection of the very
large event. Magnitude is often calculated as a distance scaled waveform amplitude. If the location is
grossly inaccurate (by tens or hundreds of kilometres), the magnitude assessed could be overestimated
by two or more orders of magnitude. Similarly, waveform corruption, such as severe electrical noise (a
large 50 or 60 Hz electrical noise) could result in a gross over-estimation of event magnitude.
 The two large events may be unidentified mine blasts.

In any case, seismic events that are much larger than expected in a population are a sign that there may be a data
integrity problem with the dataset.

Figure 61 The data is well-behaved with a linear power law relation for two orders of magnitude,
however, the largest event in the dataset is magnitude +1.5, which is almost an order of
magnitude greater than the expected largest event.

Frequency-magnitude relations can be used to identify if a significant number of mine blasts have been left in the
seismic dataset. Mine blasts are recorded by seismic systems with magnitudes in the range of 0 to +2. In
frequency-magnitude charts, leaving some blasts in the dataset has the effect of greatly increasing the number of
large events. In Figure 62(a), a few thousand mine blasts were inadvertently left in the seismic dataset. The
shape of the frequency-magnitude relation is poor, showing an abnormal number of events larger than magnitude
–0.5. When the blasts are removed from the dataset (Figure 62(b)), the frequency-magnitude relation is well-
behaved (linear from magnitude –1.8 to +1.8).

Marty Hudyma
Assistant Professor, Laurentian University 59
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

b= -0.9 b= -1.0

a b

Figure 62 Frequency-magnitude relations for the same dataset. In the chart on the left (a), a few
thousand blasts were inadvertently left in the database. On the right (b), all blasts have
been removed from the database.

Marty Hudyma
Assistant Professor, Laurentian University 60
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

REFERENCES

Amidzic, D. (2001) Energy-moment relation and its application. In Proceedings of Rockbursts and Seismicity
in Mines – RaSiM 5, Johannesburg, September 2001. ed. G. van Aswegen et al. Johannesburg: South
African Institute of Mining and Metallurgy. 509-513.
Andrieux, P.P. and Simser, B.P. (2001) Ground stability-based mine design guidelines at Brunswick Mine.
Underground Mining Methods Handbook. (Editors: W.A. Hustrulid and R.L. Bullock). Society for Mining,
Metallurgy, and Exploration Inc.: Littleton. 207-214.
Australian Centre for Geomechanics (2005) MS-RAP Version 3.1 User Documentation. 237 pages.
Basson, F.R.P. and Ras, D.J.R.M. (2005) A method to examine the time-space relationship between seismic
evens. Controlling Seismic Risk – Rockbursts and Seismicity in mines, (Editors: Y. Potvin and M.R.
Hudyma), Perth, Australian Centre for Geomechanics, pp.347-351.
Brummer, R.K. and Rorke, A.J. (1990) Case studies on large rockbursts in South African gold mines. In
Proceedings of Rockbursts and Seismicity in Mines. Minneapolis. (Editor: C. Fairhurst). Rotterdam:
A.A.Balkema. pp. 323-329.
Brummer, R.K. (1999) Simple truths about rockbursts. In Proceedings of SARES99, 2nd South African Rock
Engineering Symposium, Johannesburg, 13-15 September 1999. ed. T.O. Hagan, 6-11. Johannesburg.
Boatwright, J., and J.B. Fletcher. (1984) The partition of radiated energy between P and S waves. Bull. Seismol.
Soc. Am. 74:361-376.
Cichowicz, A., Green, R.W.E., Brink, A.v.Z, Grobler, P. and Mountfort, P.I. (1990) The space and time
variation of micro-event parameters occurring in front of an active stope. In Proceedings of Rockbursts and
Seismicity in Mines. Minneapolis. (Editor: C. Fairhurst). Rotterdam: A.A.Balkema. pp. 171-175.
Cook, N.G.W. (1976) Seismicity associated with mining. Engineering Geology. Vol. 10. p 99-122.
Cornell, C.A. (1968) Engineering seismic risk analysis. Bull. Seism. Soc. Am., 58, 1583-1606.
Duplancic, P. (2001) Characterisation of caving mechanisms through analysis of stress and seismicity.
Unpublished PhD Thesis. University of Western Australia. Perth, Australia.
Gibowicz, S.J. (1990) The mechanism of seismic events induced by mining A review. In Proceedings of
Rockbursts and Seismicity in Mines. Minneapolis. (Editor: C. Fairhurst). Rotterdam: A.A.Balkema. pp. 3-27.
Gibowicz, S.J. (1993) Keynote address: Seismic moment tnesor and the mechanism of seismic events in mines.
In Proceedings of Rockbursts and Seismicity in Mines, Kingston, August 1993. (ed R.P. Young). Rotterdam:
Balkema, 149-155.
Gibowicz, S.J., and A. Kijko (1994) An introduction to mining seismology. 1st ed. San Diego: Academic
Press.
Gibowicz, S.J. (1997) An anatomy of a seismic sequence in a deep gold mine. PAGEOPH, Vol. 150, p. 393-
414.
Gutenberg, B. and Richter, C.F. (1944) Frequency of earthquakes in California, Bulletin of the Seismological
Society of America, 34, 185-188.
Hanks, T.C. and Kanamori, H. (1979) A moment magnitude scale. Journal of Geophysical Research, 84, p.
2348-2350.
Hasegawa, H.S. (1983) Lg spectra of local earthquakes recorded by the eastern Canada telemetered network and
spectral scaling. Bulletin of the Seismological Society of America, 73 (4): 1041-1061.

Marty Hudyma
Assistant Professor, Laurentian University 61
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Hasegawa, H.S., Wetmiller, R.J. and Gendzwill, D.J. (1989) Induced seismicity in mines in Canada – an
overview. Pure and Applied Geophysics. Vol. 129, Nos 3/4, p. 423-453.
Hedley, D.G.F. (1992) Rockburst Handbook for Ontario Hardrock Mines. CANMET Special Report SP92-1E.
305 pages.
Hudyma, M.R. (1995) Seismicity at Brunswick Mining. In Proceedings Quebec Mining Association Ground
Control Colloque. Val D’Or, Quebec. 12 pages.
Hudyma, M.R., D. Milne, and D.R. Grant. (1995) Geomechanics of sill pillar mining in rockburst prone
conditions. Final report: sill pillar monitoring using conventional methods. Mining Research Directorate.
104 pages.
Hudyma, M.R., Heal, D., and Mikula, P. (2003) Seismic Monitoring in Mines – Old Technology – New
Applications. In Proceedings 1st Australasian Ground Control in Mining Conference, Sydney, p. 201-218.
Hudyma, M.R. (2004) Mining-Induced Seismicity in Underground, Mechanised, Hardrock Mines – Results of a
World Wide Survey. Australian Centre for Geomechanics Research Report.
Hudyma, M.R. (2008) Analysis and Interpretation of Clusters of Seismic Events in Mines. Unpublished PhD
thesis, University of Western Australia.
Jost, M.L. and Herrmann, R.B. (1989) A student’s guide to and review of moment tensors. Seismological
Research Letters, Volume 60, No 2. 37-57.
Kaiser, P.K., McCreath, D.R., and Tannant, D.D. (1996) Canadian Rockburst Support Handbook. 324 pages.
Kijko, A., Lasocki, S. and Retief, S.J.P. (1999) Identification of rockmass discontinuities in a cluster of seismic
event hypocenters. Safety in Mine Research Advisory Committee, GAP 622, Final Report.
Kijko, A. (1997) Seismic hazard assessment in mines. In Proceedings of Rockbursts and Seismicity in Mines,
Krakow, (ed. S. Gibowicz and S. Lasocki), 247-256.
Legge, N.B. and S.M. Spottiswoode. (1987) Fracturing and microseismicity ahead of a deep gold mine stope in
the pre-remnant stages of mining. In Proceeding of the 6th Int. Congress on Rock Mech., Montreal,
September 1987, p. 1071-1078.
Leslie, I. and Vezina, F. (2001) Seismic data analysis in underground mining operations using ESG’s Hyperion
systems. In Proceedings of the 16th Quebec Mining Association Ground Control Colloque, March 2001,
Val D'Or.
Lynch, R.A., Wuite, R., Smith, B.S. and Cichowicz, A. (2005) Microseismic monitoring of open pit slopes.
Controlling Seismic Risk - Rockbursts and Seismicity in Mines (Editors: Y. Potvin, M.R. Hudyma), Perth:
Australian Centre for Geomechanics, 581-592.
McCreary, R.G. Grant, D. and Falmagne, V. (1993) Source mechanisms, three-dimensional boundary-element
modeling, and underground observations at Ansil Mine. In Proceedings of Rockbursts and Seismicity in
Mines, Kingston, August 1993. (ed R.P. Young). Rotterdam: Balkema, 227-232.
McGarr, A. (1984) Some useful applications of seismic source mechanism studies to assessing underground
hazard. In Proceedings of the 1st International Congress on Rockbursts and Seismicity in Mines, (Editors N.
C. Gay and E. H. Wainwright), Johannesburg, South African Institute of Mining and Metallurgy, p. 199-208.
Mendecki, A.J., van Aswegen, G. and Mountfort, P. (1999) “A guide to routine seismic monitoring in mines.”
Chapter 9 in A Handbook on Rock Engineering Practice for Tabular Hard Rock Mines. A.J. Jager and J.A.
Ryder (ed.). Creda Communications, Cape Town.
Mendecki, A.J. and van Aswegen, G. (2001) Seismic monitoring in mines: selected terms and definitions. In
Proceedings of Rockbursts and Seismicity in Mines – RaSiM 5, Johannesburg, September 2001. ed. G. van
Aswegen et al, 563-570. Johannesburg: South African Institute of Mining and Metallurgy.

Marty Hudyma
Assistant Professor, Laurentian University 62
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Minney, D. Kotze, G., and van Aswegen, G. (1997) Seismic Monitoring of the Caving Process Above a
Retreating Longwall at New Denmark Colliery, South Africa. In Proceedings of Rockbursts and Seismicity
in Mines, Krakow, (ed. S. Gibowicz and S. Lasocki), pp. 125-130.
Mollison, L., Sweby, G., and Potvin, Y. (2001) Changes in mine seismicity following a mine shutdown. In
Proceedings of Advanced Rock Mechanics Practice for Underground Mines. Australian Centre for
Geomechanics. 22-23 March 2001. Perth. 13 pages.
Morrison, D.M., Swan, G. and Scholz, C.H. (1993) Chaotic behaviour and mining-induced seismicity. In
Proceedings of Rockbursts and Seismicity in Mines, Kingston, August 1993, (Editor: R.P. Young),
Rotterdam, A.A. Balkema, pp.233-237.
Nuttli, O.W. 1973. Seismic wave attenuation and magnitude relations for Eastern North America. J. Geophys.
Res., vol. 78, pp. 876-885.
Ortlepp, W.D. (1992) Invited Lecture: The design of support for the containment of rockburst damage in tunnels
– An engineering approach. Proceedings of Rock Support and Underground Construction, (editors: P.K.
Kaiser and D.R. McCreath), Rotterdam: Balkema, p. 593-609.
Owen, M. (2000) Magnitude correlations for Western Australian mines. Internal Australian Centre for
Geomechanics Report – Mine Seismicity and Rockburst Risk Management.
Richter, C.F. (1935) An instrumental earthquake magnitude scale. Bulletin of the Seismological Society of
America. Vol 25, p 1-32.
Rydelek, P.A. and Hass, L. (1994) On estimating the amount of blasts in seismic catalogs. Bulletin of the
Seismological Society of America, Vol. 84, pp. 1256-1259.
Rydelek, P.A. and Sacks, I.S. (1989) Testing the completeness of earthquake catalogues and the hypothesis of
self-similarity, Nature, Vol. 337, pp. 251-253.
Simser, B.P., Falmagne, V., Gaudreau, D., and MacDonald, T. (2003) Seismic Response to Mining at the
Brunswick Mine. CIM AGM, Montreal. 12 pages.
Spottiswoode, S.M. and McGarr, A. (1975) Source parameters of tremors in a deep-level gold mine. Bull. Seis.
Soc. Am. 65, 93-112.
Spottiswoode, S.M. (1984) Source mechanisms of mine tremors at Blyvooruitzicht Gold Mine. In Proceedings
of the 1st International Congress on Rockbursts and Seismicity in Mines, (Editors N. C. Gay and E. H.
Wainwright), Johannesburg, South African Institute of Mining and Metallurgy, p. 29-37.
Talebi, S., Mottahed, P. and Pritchard, C.J. (1997) Monitoring seismicity in some mining camps of Ontario and
Quebec. In Proceedings of Rockbursts and Seismicity in Mines, Krakow, (ed. S. Gibowicz and S. Lasocki),
117-120.
Trifu, C.I. and Urbancic, T.I. (1996) Fracture coalescence as a mechanism for earthquakes: observations based
on mining induced seismicity. Tectonophysics, 261 (1-3). 193-207.
Trifu, C.I. and Shumila, V. (2002) Reliability of seismic moment tensor inversions for induced microseismicity
at Kidd mine, Ontario, Pure and Applied Geophysics, Vol. 159, pp. 145-164.
Urbancic, T.I. (1991) Source studies of mining-induced microseismicity at Strathcona mine, Sudbury, Canada: A
spatial and temporal analysis. Unpublished PhD Thesis. Queen’s University, Kingston, Canada.
Urbancic, T.I., R.P. Young, S. Bird, and W. Bawden (1992) Microseismic source parameters and their use in
characterizing rock mass behaviour: considerations from Strathcona mine. In Proceedings of 94th Annual
General Meeting of the CIM: Rock Mechanics and Strata Control Sessions, Montreal, 26-30 April 1992. 36-
47.
Urbancic, T.I. and Trifu. C-I. (1995) Remote triggering of large seismic events in mines: evidence based on the
spatial and temporal migration of microseismic events along increasing stress gradients. In Earthquakes

Marty Hudyma
Assistant Professor, Laurentian University 63
APPLIED MINE SEISMOLOGY CONCEPTS AND TECHNIQUES JANUARY 2010

Induced by Underground Nuclear Explosions: Environment and Ecological Problems, (Editors: R. Console
and A. Nikolaev) NATO ASI Series 2, Vol.4, Springer-Verlag: Berlin, pp.353-374.
van Aswegen, G., and Butler, A.G. (1993) Applications of quantitative seismology in South African gold mine.
Rockbursts and Seismicity in Mines (editor: R.P. Young). Rotterdam: Balkema, 261-266.
van Aswegen, G. (2005) Routine seismic hazard assessment in some South African mines. Controlling Seismic
Risk - Rockbursts and Seismicity in Mines (Editors: Y. Potvin, M.R. Hudyma), Perth: Australian Centre for
Geomechanics, 437-444.
Wyss, M. and Brune, J.N. (1968) Seismic moment, stress and source dimensions for earthquakes in the
California-Nevada region. Journal of Geophysical Research, 73, 4681-4694.

Marty Hudyma
Assistant Professor, Laurentian University 64

View publication stats

You might also like