Zum Inhalt

5. Time Distribution and Seismic Hazard

  • Open Access
  • 2025
  • OriginalPaper
  • Buchkapitel
Erschienen in:

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

Dieses Kapitel befasst sich mit der Zeitverteilung und dem seismischen Risiko in Bergbauumgebungen, wobei der Schwerpunkt auf den Merkmalen seismischer Aktivität und der Anwendung statistischer Modelle zur Beurteilung und Vorhersage seismischer Risiken liegt. Es untersucht die Eigenschaften stationärer und nicht stationärer Prozesse, den Poisson-Prozess und die gestreckten Exponentialfunktionen und liefert ein umfassendes Verständnis seismischer Aktivitätsmuster. Der Text untersucht auch den Variationskoeffizienten und die proportionale Variabilität und bietet Einblicke in die Häufung und Periodizität seismischer Ereignisse. Darüber hinaus werden der Zwischenbegriff Hazard, die homogene Poisson-Verteilung sowie die Wahrscheinlichkeit von Überschreitungen und mittleren Wiederholungszeiten diskutiert und die Bedeutung dieser Konzepte für die seismische Gefahrenbewertung hervorgehoben. Das Kapitel schließt mit einer detaillierten Analyse der seismischen Gefahrenunterschiede zwischen drei Minen und liefert praktische Beispiele und Fallstudien, um die Anwendung dieser statistischen Modelle in realen Szenarien zu veranschaulichen.

5.1 Time Characteristics of Seismic Activity

Stationarity is a property of an underlying stochastic process, and not of observed data. They would be characterised by a constant mean and a constant standard deviation. Non-stationary processes have statistical properties that are deterministic functions of time. All natural processes are non-stationary, and therefore, the question is whether the underlying non-stationarity is strong enough to justify building a complex deterministic model of the process. In many cases over longer time, a simple stationary stochastic model may represent the process adequately.
A renewal process is a point process characterised by the fact that the successive inter-arrival times are distributed with the same probability density function. A Poisson process is a renewal process in which the inter-arrival times are exponentially distributed, \(f\left (t\right ) = \lambda \exp \left (-\lambda t\right )\), where the rate of events, \(\lambda \), is constant. In general, the superposition of a number of mutually independent renewal processes converges to Poisson process.
A homogeneous Poisson process results in a random series of events occurring at a constant rate, \(\lambda \), in time or at a constant density in space. It is characterised by independence, stationarity, and orderliness. Independence means that the occurrence of a given event is not influenced by the occurrence of any other event in the group—they are not correlated. Stationarity means that the rate \(\lambda \), or the underlying probability distribution, is constant over time or space, but it does not mean that all time or space intervals are equally likely.
While in a Poisson process occurrences are very irregular, the probability of an event remains constant regardless of how much time elapsed since a previous event. Orderliness precludes the possibility of multiple events at a single point in time or space or the possibility of an infinite number of events in a finite interval of time or space.
Over the short term, small seismic events in mines tend to occur in clusters in space and in time in response to rock extraction. Such processes are neither stationary nor independent. However, if a number of such processes are superimposed over a longer time and over a larger area, the outcome may not be far from Poissonian. Assumption of stationarity is very useful because it facilitates statistical analysis where probabilities are well defined as limits of frequency of occurrence.
Larger seismic events in mines are less clustered in time and in space than smaller events. Hence, the Poisson process may provide an asymptotic fit to the distribution of larger events in the time or volume mined domains. However, there are cases where larger events are clustered in time, specifically at the later stages of mining when the extraction ratio is higher and mining is deeper. While stationary models may apply to the intermediate and long term hazard assessment, over the short term one needs to resort to non-stationary models.
Note that in mines the main problem in analysing the spatial and temporal characteristics of seismicity is gaps in the data mainly due to frequent power failures. Therefore, it is advisable to test the continuity of data and, in some cases, split data into uninterrupted sections.

5.1.1 Coefficient of Variation and Proportional Variability

Coefficient of Variation
In 1896, Pearson introduced a number of statistical measures, among them the coefficient of variation, \(C_{v}\), as the ratio of the standard deviation to the mean. For the data set \(x = x_{1},...,x_{n}\),
$$\displaystyle \begin{aligned} C_{v}=\frac{s_{d}}{\overline{x}}=\sqrt{\frac{1}{n}\sum_{i=1}^{n}\left(x_{i}-\overline{x}\right)^{2}}/\left(\frac{1}{n}\sum_{i=1}^{n}x_{i}\right)=\sqrt{\frac{1}{n}\sum_{i=1}^{n}\left(\frac{x_{i}-\overline{x}}{\overline{x}}\right)^{2},}{} \end{aligned} $$
(5.1)
where \(s_{d}\) is the sample standard deviation and \(\overline {x}\) is the sample mean. If \(t_{1}\), \(t_{2},\ldots ,t_{n}\) are the times of n consecutive seismic events and \(t_{d} = t_{2}-t_{1}\),..., \(t_{n}-t_{n-1}\) are the \(n-1\) respective time differences between them, then \(C_{v}\left (t\right ) = s_{d}\left (t_{d}\right )/\overline {t}_{d}\) can be used to test for the time distribution of these events.
The time distribution of seismic activity can be either random, quasi-periodic, or clustered. If \(C_{v}\left (t\right ) \ll 1\), the process is close to periodic oscillations, and if \(0 < C_{v}\left (t\right ) < 1\), the process is quasi-periodic. For a Poissonian process, where the standard deviation is equal to the mean, \(C_{v}\left (t\right ) = 1\). For a clustered process ,\(C_{v}\left (t\right ) > 1\). For a power law distribution of inter-event times, the clustering takes an extreme form, and the mean inter-occurrence time, the standard deviation, as well as the coefficient of variation tend to infinity as the observation time increases. This form of clustering is called fractal.
Note that the fact that in a Poisson process the time intervals between events of a given size are exponentially distributed means that short intervals are more probable than long ones and events tend to cluster in small groups, separated by longer spacings. This apparent clustering does not indicate any greater dependence between the closely spaced events than between the more distant ones. In statistical physics, the stationary Poissonian process is considered as a prototype of disorder.
If the inter-event time distribution for events greater than \(\log P\) is thin tailed, e.g. the periodic, uniform, semi-Gaussian, and the stretched exponential with exponent \(q>1.0\), then the longer it has been since the last event of that size the shorter the expected time till the next one. However, if this distribution is thick tailed, e.g. the power law or the stretched exponential with exponent \(q<1.0\), then the longer it has been since the last event of that size the longer the expected time till the next one (Davis et al., 1989; Sornette & Knopoff, 1997). The case \(C_{v}\left (t\right )=1\) is memoryless; therefore, the expected time until the next event greater than or equal to \(\log P\) is independent of time.
In practice, we have a data set of n events with time span \(\Delta t = t_{n} - t_{1}\), all above a given \(\log P_{min}\), and wish to calculate the coefficient of variation of events \(C_{v}\left (t\right )\) of events with \(\log P \geq \log P_{min}\). In this case, the time difference \(\Delta t\left (\geq \log P\right ) = t_{n}\left (\geq \log P\right ) - t_{1}\left (\geq \log P\right )\) may be significantly smaller than \(\Delta t = t_{n} - t_{1}\). This underestimates the mean recurrence intervals; therefore, it overestimates seismic hazard and needs to be corrected. One way to correct is to “stretch” the observed inter-event times by \(\Delta t / \Delta t\left (\geq \log P\right )\), so that they would sum to \(\Delta t\).
The coefficient of variation is not a perfect measure of clustering, and it has its drawbacks: (1) It is problematic when the data are both positive and negative. (2) When the mean value is near zero, the coefficient of variation is sensitive to small changes in the mean, limiting its usefulness. (3) It lacks an upper bound; therefore, it is difficult to interpret. (4) It does not identify the timescales involved, so it is always useful to quote the mean as a reference. (5) It is sensitive to outliers. (6) The estimate of the coefficient of variation is a negatively biased quantity, and the bias increases as the sample size decreases. Haldane (1955) derived a small sample correction, \(\hat {C}_{v} = C_{v} \left [1+1/\left (4n\right )\right ]\), but this correction is always smaller than the expected error. Kvalseth (2016) proposed the ratio,
$$\displaystyle \begin{aligned} C_{v2}=\frac{s_{d}}{\sqrt{\overline{x^{2}}}}=\sqrt{\frac{1}{n}\sum_{i=1}^{n}\left(x_{i}-\overline{x}\right)^{2}}/\sqrt{\frac{1}{n}\sum_{i=1}^{n}x_{i}^{2}}=\sqrt{\frac{s_{d}^{2}}{s_{d}^{2}+\overline{x^{2}}}},{} \end{aligned} $$
(5.2)
called the second order coefficient of variation which has the following properties. (1) It is well defined for all real-valued variables and data. (2) Takes values between \(\left (0,1\right )\) and \(C_{v2}=0\) if \(s_{d}=0\) and \(C_{v2}=1\) only if \(\overline {x}=0\). (3) It is less affected by outliers and less sensitive to changes in the mean. For a Poissonian process, the \(C_{v}\left (t\right ) = 1\) and \(C_{v2}\left (t\right ) = \sqrt {1/2} = 0.7071\).
Proportional Variability
The proportional variability, \(P_{v}\), is quantified by comparing the numbers with each other, requiring no assumptions about central tendency or underlying statistical distributions, i.e. it is non-parametric. Like the \(C_{v}\), it provides a summary of variation where the chronology of the data is irrelevant. For a given data set of n non-negative values \(x_{i}\geq 0\), there are \(n_{k} = n\left (n-1\right )/2\) combinations of \(\left (x_{i},x_{j}\right )\) for which one can calculate the relative difference \(D\left (x_{i},x_{j}\right )\),
$$\displaystyle \begin{aligned} P_{v}=\frac{1}{n_{k}}\sum_{n_{k}}D\left(x_{i},x_{j}\right),{} \end{aligned} $$
(5.3)
where \(D\left (x_{i},x_{j}\right )=0\) if \(x_{i}=x_{j}\) and \(D\left (x_{i},x_{j}\right ) = \left |x_{i}-x_{j}\right | / \max \left (x_{i},x_{j}\right ) = 1 - \min \left (x_{i},x_{j}\right ) / \max \left (x_{i},x_{j}\right )\). The \(P_{v}\) varies between 0 and 1, it is the average percentage difference between all combinations of observed differences, therefore, unlike the \(C_{v}\), it is independent of the deviation from the mean, and it is less sensitive to outliers. It has also been shown that the \(P_{v}\) is more accurate than the coefficient of variation at estimating long term variability from short term data (Heath, 2006).
If the data is constant, \(P_{v}=0\), the time series is a simple horizontal line. The inverse of \(P_{v}\) can be interpreted as stability; therefore, \(P_{v} = 0\) would represent complete stability. If there is variability, the ordered series will be increasing and its steepness indicating the extent of variability. If the ordered series is linearly increasing, then \(P_{v}=0.5\). If \(P_{v} > 0.5\), the values in the ordered series are increasing in a non-linear way.
For a continuous probability distribution \(f\left (x\right )\) of a non-negative real variable x, the proportional variability is defined as \(P_{v} = 1 - 2\intop _{0}^{\infty }\intop _{x_{i}}^{\infty } \left (x_{i}/x_{j}\right ) f\left (x_{i}\right ) f\left (x_{j}\right ) \)\(dx_{j} dx_{i}\). For the uniform distribution, \(f\left (x\right ) = 1/\left (b-a\right )\), with the width \(w = \left (a-b\right )/2\), the standard deviation, \(s_{d} = \left (b-a\right ) / \left (2\sqrt {3}\right )\), and the coefficient of variation, \(C_{v} = \left (b-a\right ) / \left [\sqrt {3}\left (b+a\right )\right ]\). For \(w \ll \left (a+b\right )/2\), the \(P_{v}\) is very similar to \(C_{v}\), and for \(a=0\), the \(C_{v}= 1/\sqrt {3} = 0.577\), and the \(P_{v}=0.5\). For the exponential distribution, \(\lambda \exp \left (-\lambda x\right )\), the coefficient of variation \(C_{v} = 1\), and the \(P_{v} = 2 \left (1-\ln 2\right ) \simeq 0.6137\). The \(P_{v}\) can be calculated even when a mean is not defined, e.g. in case of the OE power law, \(f\left (P\right )=dF\left (P\right )/dP=\beta P_{min}^{\beta }P^{-\beta -1}\), where \(P_{v} = 1/\left (1+\beta \right )\), for any \(\beta \) and irrespective of \(P_{min}\). As expected, the lower the exponent \(\beta \), the higher the variability.
Example 1
Figure 5.1 shows the time histories of \(\log P\) in a mine over 50 months for \(\log P_{min} = -2.5\) and \(\log P_{min} = -0.5\). The coefficient of variation for 39044 events with \(\log P_{min} = -2.5\) is 2.31, which indicates relatively strong time clustering, and for 983 events with \(\log P \geq -0.5\) is 1.29, which is not far from Poissonian. The second order coefficient of variation, \(C_{v2}\), dropped from 0.92 to 0.79, and the proportional variability, \(P_{v}\), dropped only a fraction from 0.73 to 0.73. However, there are cases in mines where larger events are clustered.
Fig. 5.1
Time histories of \( \log P\) in a mine over the time span of 2130 days for two different thresholds of \( \log P_{min} = -2.5\) and \( \log P_{min}\)=\(-0.5\)
Bild vergrößern
Example 2
Figure 5.2 shows the time clustering of two short time sequences of 150 seismic events with \(-1.0 \leq \log P \leq 1.0\).
Fig. 5.2
Two time sequences of 150 seismic events simulated by the exponential distribution and the respective empirical survival functions with constant \(\tau =1.0\) for \(q = 1.0\)(left column) and by the stretched exponential distribution with \(q=0.35\)(right column)
Bild vergrößern
Both are simulated by the stretched exponential distribution, \(f\left (t\right ) = \left (q/\tau \right ) \left (t/\tau \right )^{q-1} \exp \left [-\left (t/\tau \right )^{q}\right ]\), with \(\tau =1.0\). The first one with \(q = 1.0\) that gives the exponential distribution of the inter-event times with \(C_{v}=1\), \(C_{v2}=0.71\), and \(P_{v}=0.6\). The second one with \(q=0.35\) which delivers a thick tail power law type distribution in time, therefore strongly clustered with \(C_{v}=3.02\), \(C_{v2}=0.95\), and \(P_{v}=0.83\).
The empirical survival function, \(\Pr \left (\geq t\right )\), for the data set that simulates the exponential distribution is linear for the entire domain of inter-event times, and probabilities of outliers are low. However, the clustered data set with \(C_{v}=3.02\) delivers a highly non-linear survival function with much higher probabilities of extreme inter-event times.

5.2 Intermediate Term Hazard

The mean recurrence interval between events above a certain potency estimated over the period of time \(\Delta t\) is \(\bar {t}\left (\geq P\right ) = \Delta t / N\left (\geq P\right )\), where \(\Delta t\) is the time span of data used and \(N\left (\geq P\right )\) is the number of events not smaller than P over that time. Seismic activity rate for these events is then \(1/\bar {t}\left (\geq P\right )\). The mean volume mined between events above a certain size during extraction of \(V_{m}\) volume of rock is \(\bar {V}\left (\geq P\right ) = V_{m} / N\left (\geq P\right )\). The number of events, \(N\left (\geq P\right )\), per unit of volume of rock extraction is \(1 / \bar {V}\left (\geq P\right )\). Note that the terms “mean inter-event time”, “mean recurrence interval” may be deceptive since in many cases the dispersion from the mean, as measured by the standard deviation, is comparable with the mean value. If the standard deviation of the observed recurrence intervals is less than 50% of the mean recurrence interval, then the seismic behaviour may be assumed to be periodic rather than episodic, and calculated probability estimates can be considered reasonable. For a very short time interval into the future, \(\Delta T < \bar {t}\left (\geq P\right )\), the probability of having an event with potency not smaller than P can be estimated as \(\Pr (\geq P) \simeq \Delta T / \bar {t}\left (\geq P\right )\). The calculated recurrence interval or the mean volume mined that stretch well beyond the span of the data set \(\Delta t\) or \(V_{m}\) should be treated with caution.

5.2.1 Homogeneous Poisson Distribution

The Poisson distribution can be used to forecast the number of random events in a future time, \(\Delta T\). There are two random variables that arise from a Poisson process: (1) the number of events in a given time interval, which is a discrete variable, (2) the time until the occurrence of the first event, which is a continuous variable. Having a single event in the time interval \(\Delta t\), the probability of finding this event in the sub-interval \(\Delta T = t_{2} - t_{1}\) is \(\Delta T/\Delta t\). Having n events in \(\Delta t\), the activity rate is \(\lambda _{t} = n/\Delta t\), the recurrence time \(\bar {t} = \Delta t / n\), and the expected number of events in \(\Delta T\) is \(\Lambda \left (\Delta T\right ) = n \left (\Delta T/\Delta t\right ) = \lambda _{t}\Delta T\). The probability that exactly \(N < n\) events can be found in the interval \(\Delta T\) is given by the binomial distribution,
$$\displaystyle \begin{aligned} \Pr\left(N,\Delta T\right)=n!\left[\Lambda\left(\Delta T\right)/n\right]^{N}\left\{ 1-\Lambda\left(\Delta T\right)/n\right\} ^{n-N}/\left[N!\left(n-N\right)!\right].{} \end{aligned} $$
(5.4)
The mean of the binomial distribution is \(n\Delta T/\Delta t\), and the variance is \(n\Delta T/\Delta t \left (1-\Delta T/\Delta t\right )\), i.e. the variance is less than the mean. For large n and small \(\Delta T/\Delta t\), the binomial distribution can be approximated by Poisson distribution,
$$\displaystyle \begin{aligned} \Pr\left(N,\Delta T\right)=\frac{\left[\Lambda\left(\Delta T\right)\right]^{N}}{N!}\exp\left[-\Lambda\left(\Delta T\right)\right]=\left(\frac{\Delta T}{\bar{t}}\right)^{N}\frac{1}{N!}\exp\left(-\Delta T/\bar{t}\right),{} \end{aligned} $$
(5.5)
with the mean value and the variance equal to \(\lambda _{t} \Delta T = n\Delta T/\Delta t\) and the coefficient of variation \(C_{v} = 1/\left (\lambda _{t}\Delta T\right )\). Note that \(\Pr \left (N=1,\bar {t}\right ) = 0.37\).
The probability that there will be no events in \(\Delta T\) is \(\Pr \left (N=0,\Delta T\right ) = \exp \left (-n\Delta T/\Delta t\right )\), and the probability that there will be at least one event is \(F\left (\Delta T\right ) = \Pr \left (N\geq 1,\Delta T\right ) = 1 - \exp \left (-n\Delta T/\Delta t\right )\), which is the exponential distribution with the mean \(1/\lambda _{t}\) and the variance \(1/\lambda _{t}^{2}\), and therefore, the coefficient of variation \(C_{v} = 1\). If \(n \left (\Delta T/\Delta t\right ) < 0.1\), then \(\Pr \left (N\geq 1,\Delta T\right ) \simeq n \left (\Delta T/\Delta t\right )\). Note that \(\Pr \left (N\geq 1,\bar {t}\right ) = 0.63\), see Fig. 5.3.
Fig. 5.3
Left. Poisson probabilities of having exactly 1 (green solid line) and exactly 5 (green dotted line) events as a function of waiting time \(\Delta T\). Exponential probabilities of having at least one event (blue solid line) and no events (blue dotted line). Right. Probability density function of an exponential distribution (blue) and probabilities that an event will occur in time interval \( \left (0,1/\lambda \right )\) or \( \left (1/\lambda ,\infty \right )\)
Bild vergrößern
The exponential distribution describes the length of time between events, or the inter-event time distribution, and it does not depend on time. The assumption of stationary and independent increments means that at any point the process probabilistically restarts itself, i.e. it has no memory. The probability density function of an exponential distribution is \(f\left (\Delta T\right ) = \lambda \exp \left (-\lambda \Delta T\right )\), and the probability that an event will occur in time interval \(\left (0,1/\lambda \right )\) is \(\Pr \left (0,1/\lambda \right ) = \intop _{0}^{1/\lambda }\lambda \exp \left (-\lambda \Delta T\right )d\left (\Delta T\right ) = 0.632\), and the probability that it will occur in the time interval \(\left (1/\lambda ,\infty \right )\) is \(\Pr \left (1/\lambda ,\infty \right ) = 0.368\), see Fig. 5.3.

5.2.2 Probabilities of Exceedance and Mean Recurrence Times

If a set of n random variables, \(X_{1}\), \(X_{2}\), ..., \(X_{n}\), where \(X_{1} = P_{1}\), \(X_{2} = P_{2}\), ... \(X_{n} = P_{n}\) is one possible realisation of each of them, is drawn from the same distribution function \(F\left (P\right ) = \Pr \left (\leq P\right )\), then the distribution of the maximum value \(P_{max} = \max \left \{ P_{1},P_{2},...,P_{n}\right \} \) is given by \(\Pr \left (\leq P\right ) = \Pr \left (P_{1}\leq P,P_{2}\leq P,...,P_{n}\leq P\right ) = \prod _{j=1}^{n}\Pr \left (\leq P_{j}\right ) = \left [F\left (P\right )\right ]^{n}\).
If the occurrence of larger events in time \(\Delta t\) is ruled by a homogeneous Poisson process with activity rate \(\lambda _{t} = n/\Delta t\), then the probability of having exactly N events within time \(\Delta T\) is
$$\displaystyle \begin{aligned} \Pr\left(N,\Delta T\right)=\left[\Lambda\left(\Delta T\right)\right]^{N}\exp\left\{ -\Lambda\left(\Delta T\right)\right\} /N!,{} \end{aligned} $$
(5.6)
where \(\Lambda \left (\Delta T\right ) = \lambda _{t}\Delta T\) is the expected number of events in \(\Delta T\). The probability that all \(P_{j}\) within \(\Delta T\) are smaller than or equal to some large P is
$$\displaystyle \begin{aligned} \Pr\left(\leq P,\Delta T\right)=\sum_{N=0}^{\infty}\left\{ \left[F\left(P\right)\right]^{N}\left[\Lambda\left(\Delta T\right)\right]^{N}\exp\left[-\Lambda\left(\Delta T\right)\right]/N!\right\} , \end{aligned}$$
which gives \(\Pr \left (\leq P,\Delta T\right ) = \exp \left \{ -\Lambda \left (\Delta T\right )\left [1-F\left (P\right )\right ]\right \} \), or \(\Pr \left (\leq P,\Delta T\right ) = \exp \left [-\Lambda \left (\Delta T\right )\Pr \left (\geq P\right )\right ]\), and the probability of having at least one event \(\geq P\) within \(\Delta T\),
$$\displaystyle \begin{aligned} \Pr\left(\geq P,\Delta T\right)=1-\exp\left[-\Lambda\left(\Delta T\right)\Pr\left(\geq P\right)\right]=1-\exp\left[-\left(\Delta T/\Delta t\right)N\left(\geq P\right)\right],{} \end{aligned} $$
(5.7)
where \(\Pr \left (\geq P\right ) = N\left (\geq P\right ) / n\). For the UT power law, the Eq. (5.7) gives
$$\displaystyle \begin{aligned} \Pr\left(\geq P,\Delta T\right)=1-\exp\left[-\alpha\left(\Delta T/\Delta t\right)\left(P^{-\beta}-P_{max}^{-\beta}\right)\right].{} \end{aligned} $$
(5.8)
Epstein and Lomnitz (1966) showed that the OE power law, \(N\left (\geq P\right ) = \alpha P^{-\beta }\), gives \(\Pr \left (\leq P,\Delta T\right )=\exp \left [-\alpha \left (\Delta T/\Delta t\right )\exp \left (-\beta m\right )\right ]\), where \(m = \ln P\), which is the first Gumbel distribution of the extreme values (see also Shakal & Willis, 1972 or Kijko, 1982). Note that replacing \(\Delta T\) with the recurrence time \(\bar {t}\left (\geq P\right ) = \Delta t / N\left (\geq P\right )\) gives \(\Pr \left [\geq P,\bar {t}\left (\geq P\right )\right ] = 0.63\), regardless of P. The mean recurrence time can also be presented as
$$\displaystyle \begin{aligned} \bar{t}\left(\geq P\right)=\frac{\Delta t}{N\left(\geq P\right)}=-\frac{\Delta T}{\ln\left[1-\Pr\left(\geq P,\Delta T\right)\right]},{} \end{aligned} $$
(5.9)
where \(\Pr \left (\geq P,\Delta T\right )\) is the probability of exceedance, and \(\Delta T\) here is called the exposure time. Most seismic building codes are based on the following performance levels that define the ability of a structure to sustain its main functions during and after earthquakes of different strengths. (1) Operational Limit: 50% probability of exceedance over 50 years, which is considered frequent, that according to Eq. (5.9) gives \(\bar {t}\left (\geq P\right ) = 72\) years, (2) Immediate Occupancy: 20% in 50 years (occasional) that gives \(\bar {t}\left (\geq P\right ) = 225\) years, (3) Life Safety: 10% in 50 years (rare) that gives \(\bar {t}\left (\geq P\right ) = 475\) years, and (4) Collapse Prevention: 2% in 50 years (very rare) that gives \(\bar {t}\left (\geq P\right ) = 2475\) years. In general, one expects better structural performances for frequent events of lower intensities, and one can accept higher damage for very rare large events. Building codes also adjust seismic input level to the importance of a given structure, e.g. by increasing the exposure time \(\Delta T\) in Eq. (5.9) that leads to an increase of the mean recurrence time. For many structures in mines, the exposure times \(\Delta T\) are shorter. For example, 50% probability of exceedance over 10 years would give \(\bar {t}\left (\geq P\right ) = 14.4\) years, 20% in 10 years would give \(\bar {t}\left (\geq P\right ) = 44.8\) years, and 10% in 10 years would give \(\bar {t}\left (\geq P\right ) = 95\) years.
For a Poisson process, the probability of having N occurrences over a time window equal to the recurrence time, \(\Delta T = \bar {t}\), is \(\Pr \left (N\right ) = \left (1/N!\right )\exp \left (-1\right )\), which gives a probability of 36.8% that such an event will occur once, 18.4% that it will occur twice and 6.1% for three times.
Equation (5.9) can also be presented in terms of potency as a function of recurrence times, which for the upper truncated power law, \(N\left (\geq P\right ) = \alpha \left (P^{-\beta }-P_{max}^{-\beta }\right )\), gives
$$\displaystyle \begin{aligned} P\left(\bar{t}\geq P\right)=\left[\frac{1}{\alpha}\frac{\Delta t}{\bar{t}\left(\geq P\right)}+P_{max}^{-\beta}\right]^{-1/\beta},{} \end{aligned} $$
(5.10)
and for the OET distribution \(\overline {t}\left (\geq P\right ) = \Delta t/ \left [N\left (\geq P_{min}\right )\Pr \left (\geq P\right )_{OET}\right ]\).
These equations can be used to calculate the expected potencies over the selected recurrence times and to estimate their probabilities (Eq. 5.7) as a function of \(\Delta T\), as mining progresses, to monitor longer term hazard. In this case, \(P_{max}\) should be taken as the latest estimate of the maximum expected potency. Figure 5.4 left shows the probabilities of exceedance versus \(\log E\) for MineC over 90, 180, and 360 days given by Eq. (5.8), and Fig. 5.4 right shows the probabilities of exceedance of \(\log E\) with the recurrence times of 1, 2, 3, and 5 years for the same data set calculated with Eq. (5.10).
Fig. 5.4
Probabilities of exceedance for MineC derived from Eq. (5.8) for \( \log E\)(left) and the probabilities of exceedance of \( \log E\) with the recurrence times of 1, 2, 3, and 5 year derived from equation 5.10
Bild vergrößern
Note that the term “the mean recurrence time” or “the return period” as it is frequently called may be misleading unless it also gives the standard deviation around the mean. In practice, the Poisson model is justified if this standard deviation is not far from the mean.
In mines where the volume of rock extraction is not a linear function of time seismic hazard can be expressed in terms of the volume of rock extracted, or volume mined \(V_{m}\). Then, the probability that there will be at least one event \(\geq P\) while extracting an additional volume of rock \(\Delta V_{m}\) is
$$\displaystyle \begin{aligned} \Pr\left(\geq P,\Delta V_{m}\right)=1-\exp\left[-\Lambda\left(\Delta V_{m}\right)\Pr\left(\geq P\right)\right]=1-\exp\left[-\frac{\Delta V_{m}}{V_{m}}N\left(\geq P\right)\right],{} \end{aligned} $$
(5.11)
where \(\Lambda \left (\Delta V_{m}\right ) = n \left (\Delta V_{m}/V_{m}\right ) = \lambda _{m}\Delta V_{m}\) is the expected number of events in \(\Delta V_{m}\). Since \(\bar {V}_{m}\left (\geq P\right ) = V_{m} / N\left (\geq P\right )\), the Eq. (5.11) gives the ratio \(\bar {V}_{m}\left (\geq P\right ) / \Delta V_{m} = -1 / \ln \left [1-\Pr \left (\geq P,\Delta V_{m}\right )\right ]\). For example, if \(\Pr \left (\geq P,\Delta V_{m}\right ) = 0.05\), we can mine on average 19.5 times \(\Delta V_{m}\) to generate seismic event \(\geq P\). For the UT power law, the Eq. (5.11) gives
$$\displaystyle \begin{aligned} \Pr\left(\geq P,\Delta V_{m}\right)=1-\exp\left[-\Delta V_{m}\frac{\alpha}{V_{m}}\left(P^{-\beta}-P_{max}^{-\beta}\right)\right],{} \end{aligned} $$
(5.12)
which shows that if \(\beta \) is constant and \(\alpha \) is proportional to \(V_{m}\), then seismic hazard can be controlled by the rate of mining.
Limited Data and Uncertain Activity Rate
For limited data and/or for an uncertain activity rate, one can treat the activity rate \(\lambda _{t}\) as a random variable that can be accounted for by taking the integral \(\intop _{0}^{\infty } \Pr \left (\geq P,\Delta T\right ) f\left (\lambda _{t}\right ) d\lambda _{t}\), where \(f\left (\lambda _{t}\right ) = \Delta t \left (\lambda _{t}\Delta t\right )^{n} \exp \left (-\lambda _{t}\Delta t\right ) /n!\) is the probability density function of the activity rate, which gives
$$\displaystyle \begin{aligned} \Pr\left(\geq P,\Delta T\right)=1-\left[1+\left(\Delta T/\Delta t\right)\Pr\left(\geq P\right)\right]^{-n-1}.{} \end{aligned} $$
(5.13)
The ratio of expressions (5.13) to (5.7) is approximately \(1 + 1/n\), and the probability premium \(1/n\) drops quickly with n, and for \(n \geq \) 100 is less than 1% (McGuire, 1977). Equation (5.13) can also be expressed in the volume mined domain, \(\Pr \left (\geq P,\Delta V_{m}\right ) = 1 - \left [1+\left (\Delta V_{m}/V_{m}\right )\Pr \left (\geq P\right )\right ]^{-n-1}\).
Hazard Trend
In mines, the parameters \(\alpha \) and \(\beta \) of the potency frequency distribution are not always constant or fluctuating in time as may be assumed in crustal seismology. As mining progresses, the overall stiffness of the rock mass is being degraded, and therefore, the parameters \(\alpha \beta \) and \(P_{max}\) are expected to be a function of the effective volume mined, \(V_{meff}\). The data shows that the exponent \(\beta \) decreases as the effective volume mined, i.e. extraction ratio of the ore body, increases. \(P_{max}\) also tends to increase as mining progresses. In most cases \(\alpha \) has been shown to be proportional to \(V_{meff}\), though at different rates in different mining scenarios. However, there are examples where \(\alpha \) increases with \(V_{m}\) in a non-linear way, which is indicative of increasing hazard. Figure 5.5 shows the behaviour of \(\alpha \) and \(\beta \) versus \(V_{m}\) in two mines.
Fig. 5.5
Parameters \(\alpha \) (triangles) and \(\beta \) (circles) versus the effective volume mined for two mines
Bild vergrößern
By extrapolating parameters \(\alpha \) and \(\beta \) beyond the observed volume mined or the observed time span, one can estimate future hazard for different mining scenarios from equation,
$$\displaystyle \begin{aligned} \Pr\left(\geq P,\Delta V_{m}\right)=1-\left[1+\Delta V{}_{m}/V_{m}\Pr\left[\geq P,\beta\left(\Delta V_{m}\right)\right]\right]^{-n\left[\alpha\left(\Delta V_{m}\right),\beta\left(\Delta V_{m}\right)\right]-1},{} \end{aligned} $$
(5.14)
where n is the expected number of events as a function of the future \(\alpha \) and \(\beta \) (Mendecki, 2008).

5.3 Empirical Probabilities from the Observed Recurrence Times

Given the latest n observed recurrence intervals of events \(\geq \log P\), of which \(n_{\Delta T}\) are smaller than or equal to \(\Delta T\), one can estimate the empirical probability, \(\Pr (\geq \log P,\Delta T)_{E}\), that a given volume will produce an event \(\geq \log P\) within time \(\Delta T\) after the preceding event of this size,
$$\displaystyle \begin{aligned} \Pr(\geq\log P,\Delta T)_{E}=p=\frac{n_{\Delta T}+1}{n+2},\qquad s_{d}\left(p\right)=\pm2\sqrt{\frac{p(1-p)}{n+3}},{} \end{aligned} $$
(5.15)
where \(s_{d}\left (p\right )\) is the uncertainty of the estimate. Note that the standard deviation, \(sd\left (p\right )\), decreases slowly with increasing n, and if the number of observed intervals n is not much greater than 10, the probability generally will not be determined better than \(\pm 0.2\) (Savage, 1994). For a given n, the standard deviation, or the uncertainty, is maximum at \(p=0.5\), and it decreases symmetrically to zero as \(p \rightarrow 0\) and \(p \rightarrow 1.0\). Equation (5.15) is derived under the assumption that the probability of having an event of certain size within a specified time interval following the preceding event of that size is the same for all event cycles. Empirical probabilities cannot extend to the largest few events and cannot be extrapolated. However, they can be used to test results if there is a sufficient number of the observed recurrence intervals between intermediate and larger events.
Given the latest n observed inter-event volume mined of events \(\geq \log P\), of which \(n_{\Delta V_{m}}\) are smaller than or equal to \(\Delta V_{m}\), one can estimate the empirical probabilities in the volume mined domain, \(\Pr \left (\geq \log P,\Delta V_{m}\right )_{E} = \left (n_{\Delta V_{m}}\right ) / \left (n+2\right )\).
Table 5.1 lists the empirical probabilities \(\Pr (\log P\geq 1.2,t_{1}+\Delta T)_{E}\), calculated over the periods of one day, one week and one month, having 15 events with \(\log P\geq 1.2\) in the data set with the following 14 recurrence intervals quoted here in hours: 118, 542, 265, 22, 587, 116, 56, 110, 11, 282, 95, 73, 235, 1. The last event with \(\log P\geq 1.2\) occurred at \(t_{1} =\) 14h14’ on 25 December 2014. The empirical probabilities are listed in Table 5.1.
Table 5.1
Empirical probabilities
\(\Pr (\log P\geq 1.2, \Delta T=1\,\mbox{day})_{E}\)
\(\Pr (\log P\geq 1.2, \Delta T=7\,\mbox{days})_{E}\)
\(\Pr (\log P\geq 1.2, \Delta T=30\,\mbox{days})_{E}\)
\(0.25\,(\pm 0.21)\)
\(0.62\,(\pm 0.23)\)
\(0.94\,(\pm 0.12)\)
Figure 5.6 left shows an example of the potency frequency data where for the most part it fits the upper truncated power law reasonably well, although it deviates from the 95% confidence limit in the potency range \(0.8 \leq \log P \leq 1.4\), mostly at \(1.0 \leq \log P\leq 1.2\). Figure 5.6 right shows the empirical probabilities derived from Eq. (5.15). It shows that the UT power law model underestimates seismic hazard in that potency range for the time period of the next 70 days. However, beyond this time range, the UT power law model predicts reasonably well. Note that the empirical probabilities are conditional in nature, i.e. it gives the probability within time \(\Delta T\), after the preceding event of this size.
Fig. 5.6
Potency frequency plot (left) and the empirical probabilities (right) for \( \log P \geq 0.8\), marked by the black step line, \( \log P \geq 1.0\), green, \( \log P \geq 1.2\), blue, and \( \log P \geq 1.4\), red. The smooth dashed lines are the respective probabilities estimated by Eq. (5.8) for the same data set
Bild vergrößern

5.4 Example: Seismic Hazard Difference Between Three Mines

Data Sets
Data sets from three mines: A, B, and C, were collected over the same two year period \(\Delta t =\) 678 days, all related to tabular mining with principal vertical stresses but with different geological structures, mining layouts, extraction ratios, depths, and rates of mining.
MineA practiced long-wall mining of highly extracted tabular reef, see Fig. 5.7 left column, and MineB applied the sequential grid method, see Fig. 5.7 centre. For a review of these mining methods, see Vieira et al. (2001). In both cases, the rock extraction took place at practically the same depth of 3300 m. A simple numerical elastic model shows that due to the higher extraction ratio the mean vertical stress calculated over the un-mined areas in MineA is 1.7 times higher than in MineB. MineC is related to the scattered mining imposed by the presence of larger faults at an average depth of 1755 m, see Fig. 5.7 right column.
Fig. 5.7
Mine layouts in plan and in section of MineA (left column), MineB (centre), and MineC (right)
Bild vergrößern
The depth of mining, the volume mined, the rate of mining, the approximate extraction ratio are listed in Table 5.2. The relative hazard rating from (1) the highest to (3) the lowest, imposed by the author for each parameter, is also quoted in parentheses, where applicable. Here we assumed that seismic hazard scales positively with the depth of mining, the deepest being MineA, the volume of rock extraction, the highest being MineB, the rate of mining, also MineB and with the extraction ratio, the highest being MineA. While these criteria are sensible, they are not sufficient to rate these three mines regarding seismic hazard.
Table 5.2
Mining factors with hazard rating in parentheses
Parameter
MineA
MineB
MineC
Time of observation, \(\Delta t\), days
678
678
678
Weighted depth of extraction, m
3294 (1)
3287 (2)
1755 (3)
Volume mined, \(V_{m}\), m\(^{3}\)
207688 (3)
673209 (1)
386158 (2)
Rate of mining, m\(^{3}\)/day
306 (3)
993 (1)
570 (2)
Approximate extraction ratio, %
80 (1)
70 (2)
60 (3)
Layout, Sequence of Mining, and Relative Hazard
To gain insight into the influence of the mine layout on seismic hazard, their respective fractal dimensions were estimated. Fractals appear similar at any scale of observation. In mathematical terms, fractal objects exhibit fractional dimensionality, that is, they are neither lines, nor surfaces or volumes. Their dimension falls in between the classical dimensions of Euclidean geometry. An object is fractal when its length L is a function of the length \(\lambda \) of the measuring device, \(L\sim \lambda ^{1-D}\), where D is the fractal dimension (Mandelbrot, 1967, 1975). If \(N\left (\lambda \right )\) is the number of cubes of size \(\lambda \) needed to cover the object, then the box counting fractal dimension of an object can be estimated by \(D = \ln N\left (\lambda \right ) / \ln \left (1/\lambda \right )\) (Barnsley, 1988). Fractal dimension increases with the degree of irregularity, or raggedness of the object. The lowest fractal dimension of mine layout, \(D_{ml} = 1.67\), is associated with long-wall mining in MineA, followed by sequential grid, \(D_{ml} =\) 1.68 in MineB, and the roughest is the scattered mining imposed by the presence of geological structures \(D_{ml} = 1.81\), in MineC. This sequence could easily be inferred just by looking at the smoothness of lines in Fig. 5.7.
Having coordinates and the dates and times of panel, extraction lines between consecutively extracted panels were connected, and the fractal dimension of such a spatial and temporal graph was calculated, \(D_{ste}\). The smoothest sequence of mining, \(D_{ste} = 1.84\), is associated with MineA, then MineB with \(D_{ste} = 1.92\), and the roughest by MineC with \(D_{ste} = 2.20\). By connecting lines between consecutive seismic events, an image of the sequence of seismic activity was created, and then its fractal dimension calculated. This exercise was limited to the \(\left (x,y,t\right )\) domain since the tabular ore bodies imposed a flat distribution of seismic stations and made the z-coordinates less reliable, specifically at the fringes of the network. Results and the relative hazard ratings are given in Table 5.3. They show that seismic activity does not follow mining exactly. One may speculate that the high stress and high rate of mining associated with data set B aligned most events with the excavation faces lowering the fractal dimension of their spatial distribution, but this has not been tested numerically.
Table 5.3
Mining factors, with hazard rating in parentheses
Parameter
MineA
MineB
MineC
\(D_{ml}\) of mine layout
1.67 (1)
1.68 (2)
1.71 (3)
\(D_{sxy}\) of epicentres
1.60 (2)
1.49 (1)
1.61 (3)
\(D_{ste}\) of space-time extraction
1.84 (1)
1.92 (2)
2.20 (3)
\(D_{sxyt}\) of epicentres and time
1.84 (3)
1.71 (1)
1.83 (2)
There is also a negative correlation between the fractal dimension of the epicentres and time of events, \(D_{sxyt}\), and \(\beta \), see Tables 5.3 and 5.5. There are reports stating the positive correlation for earthquakes, but mainly for single fracturing or single fault processes, e.g. Aki (1981), King (1983), Wyss et al. (2004), Chen et al. (2006). Hirata (1989) reported a negative correlation due to different fault systems and Henderson et al. (1999) for induced seismicity where they show a negative correlation for high loading rates and a positive correlation for slowly loaded systems. Amitrano (2003) also reported negative correlation stating that diffused damage is associated with low \(\beta \), whereas localised damage is associated with high \(\beta \). The data set B has the highest loading rate of all three and the lowest fractal dimension, followed by data set C, and the lowest loading rate and the highest fractal dimension is associated with data set A.
Observed Seismicity and Relative Hazard
Table 5.4 lists the observed seismic parameters and the relative hazard rating from the highest (1) to the lowest (3), imposed by the author for each parameter.
Table 5.4
Observed seismic parameters with hazard rating in parentheses
Parameter
MineA
MineB
MineC
\(\log P_{maxo}\); \(\log P_{maxo-1}\)
3.37; 2.97 (1)
2.70; 2.59 (2)
2.51; 2.43 (3)
\(N_{obs} ( \log P \geq 2.0 )\)
23 (1)
23 (1)
13 (2)
\(\bar {t} = \Delta t / N_{obs} ( \log P \geq 2.0 )\), days
29.49 (1)
29.49 (1)
52.17 (3)
\(\bar {V}_{m} = V_{m} / N_{obs} ( \log P \geq 2.0 )\), m\(^{3}\)
9.03 (1)
29.27 (2)
29.70 (3)
\(\log E_{maxo}\); \(\log E_{maxo-1}\)
9.53; 9.49 (1)
8.68; 8.63 (2)
7.75; 7.58 (3)
\(N_{obs} ( \log E \geq 7.0 )\)
98 (1)
89 (2)
12 (3)
\(\bar {t} = \Delta t / N_{obs} ( \log E \geq 7.0 )\), days
6.92 (1)
7.61 (2)
56.51 (3)
\(\bar {V}_{m}=V_{m}/N_{obs}\left (\log E\geq 7.0\right )\), m\(^{3}\)
2119.3 (1)
7564.1 (2)
32179.87 (3)
Table 5.5
Size distribution parameters for potency and energy with hazard rating in parentheses
Parameter
MineA
MineB
MineC
\(\log \alpha _{EUT}\)
6.4476 (2)
7.8055 (1)
5.6835 (3)
\(\beta _{EUT}\)
0.6517 (2)
0.832 (3)
0.6254 (1)
\(\log E_{nrb}\); \(\log E_{max}\)
10.167; 10.917 (1)
9.058; 9.524 (2)
8.257; 8.812 (3)
\(\Pr \left (\log E\geq 7;\:\Delta T=\text{30}\:\text{days}\right )\)
0.1966 (2)
0.984 (1)
0.527 (3)
\(\Pr \left (\log E\geq 8;\:\Delta T=\text{30}\:\text{days}\right )\)
0.527 (1)
0.418 (2)
0.063 (3)
\(\Pr \left (\log E\geq 7;\:\Delta V_{m}=\text{10}^{4}\:\text{m}^{3}\right )\)
0.975 (1)
0.752 (2)
0.354 (3)
\(\Pr \left (\log E\geq 8;\:\Delta V_{m}=\text{10}^{4}\:\text{m}^{3}\right )\)
0.557 (1)
0.166 (2)
0.038 (3)
Figure 5.8 shows the cumulative graphs of the volume mined, seismic potency, and energy vs. time for all three mines.
Fig. 5.8
All cumulative for MineA, MineB, and MineC: volume mined versus time, potency release versus time, and seismic energy versus time
Bild vergrößern
Figure 5.9 shows the cumulative potency and seismic energy versus volume mined. The potency release and seismic energy versus volume mined plots are evident, and both rate the hazard potential, i.e. assuming the same mining rate, for the three data sets clearly: The highest is MineA (1), the second highest is MineB (2), and the lowest of the three is MineC (3). Note that MineA mined at much lower rate to mitigate seismic hazard.
Fig. 5.9
Cumulative potency (left) and seismic energy (right) vs. volume mined
Bild vergrößern
Size Distribution Parameters and Seismic Hazard
Seismic hazard rating is a useful but a rudimentary scale, and it does not differentiate across the size distribution of seismic events. It is frequently the case that one mining area may have higher hazard up to certain event size and lower above it, and if the crossover is at the size that is potentially damaging, it may change seismic risk because the events below the threshold are more frequent.
Figure 5.10 top row shows the history of records for all three mines where \(\log E_{max}\), marked by black line, is the upper limit of the next record breaking \(\log E\), \(\log E_{nrb}\), marked by red line, is the expected next record estimated by the upper truncated distribution (UT). The green line shows the level of the expected next record estimated by the open-ended tapered distribution (OET).
Fig. 5.10
History of records for MineA, MineB, and MineC with the estimated range of \( \log P_{nrb}\)(top row) and energy frequency plots (bottom row)
Bild vergrößern
Figure 5.10 bottom row shows the cumulative number of unbinned data versus \(\log E\) for all three mines where colour indicates the time of the event, with 4 fits: OE shown in blue, the UT assuming that the upper limit of the next record breaking \(\log E = \log E_{max}\) in black, the UT assuming that the upper limit of the next record breaking \(\log E = \log E_{nrb}\) in red, and OET in green. The light grey vertical spikes below the data illustrate what would be the empirical pdf if the data was binned with bin size 0.05. The dashed grey lines on both sides of the UT power law fit indicate 95% confidence limits.
The data set for MineB and MineC shows the frequency of the largest observed events deviated from the predicted UT model; therefore, one can infer that the size distribution hazard may be contained or controlled. The frequency of the largest observed events for MineA is close to the OE relation, indicating a potential for even larger events.
In general, seismic hazard should scale positively with \(\alpha \) and negatively with \(\beta \); however, Table 5.5 shows that \(\alpha \) indicates the highest hazard for MineB and \(\beta \) for MineC. The expected next record breaking in energy, \(\log E_{nrb}\), and its upper limit, \(\log E_{max}\), rate seismic hazard correctly.
Figure 5.11 left shows seismic hazard estimated by Eq. (5.8) in terms of \(\log E\) for all three mines in the time domain for \(\Delta T = 30\) days, assuming their current rate of mining. Seismic hazard at MineC is clearly the lowest. However, there is the crossover point at \(\log E=7.4\) between MineA and MineB below which seismic hazard for MineB is higher. Figure 5.11 right shows seismic hazard potential, estimated by Eq. (5.12), assuming all three mines extract the same volume of rock \(\Delta V_{m} = 10000\ \mathrm {m}^{3}\), otherwise for the same set of size distribution parameters. Clearly, hazard potential at MineA is the highest of all three mines, and the difference between MineA and MineB is significant. Therefore, slowing down rate of mining in MineB would lower seismic hazard considerably.
Fig. 5.11
Seismic hazard for the three mines in the time domain (left) and in the volume mined domain (right)
Bild vergrößern

5.5 Example: Seismic Hazard Difference in Time

In this example, we compare seismic hazard characteristics of two mostly overlapping seismic and rock extraction data sets selected from the same mine:
1.
The DataSet1 starts on 07 September 2007 and ends on 13 June 2012.
 
2.
The DataSet2 starts on 07 September 2007 and ends on 07 July 2013, so it is DataSet1 plus an additional 389 days of mining.
 
Seismic Activity and Production During DataSet1
This data set starts on 07 September 2007 and ends on 13 June 2012 just after a \(\log P=2.24 \left (m=2.41\right )\) event. It spans 1741 days and includes 2281 events with \(\log P\geq -1.0 \left (m\geq 0.25\right )\), which gives the rate of 1.31 events/day. These events delivered \(\varSigma P = 1630.1\) m\(^{3}\) of seismic potency at the rate of 0.9363 m\(^{3}\)/day. The largest event in DataSet1 has \(\log P=2.24 \left (m=2.4\right )\), and there are 20 events with \(\log P\geq 1.0 \left (m\geq 1.59\right )\). The mean recurrence interval, \(\bar {t}\left (\log P\geq -1.0\right ) = 0.7639\) days with the standard deviation of \(1.133\) days, which gives the coefficient of variation \(C_{v}\left (\log P\geq -1.0\right ) = 1.48\). The mean recurrence interval, \(\bar {t}\left (\log P\geq 1.0\right ) = 91.634\) days with the standard deviation of \(110.75\) days, which gives the coefficient of variation \(C_{v}\left (\log P\geq 1.0\right ) = 1.21\), which indicates that larger events are less clustered in time than small events. The vertical black line in Fig. 5.12 marks the time of the second largest event during that time with \(\log P=1.82\).
Fig. 5.12
Cumulative number of events (left), cumulative potency, and cumulative volume mined vs. time (right) for DataSet1
Bild vergrößern
During this time, the mine did 1071 blasts and extracted 1092912.9 m\(^{3}\) of rock that gives the average extraction rate of 627.735 m\(^{3}\)/day. The ratio \(\Sigma V_{m}/\Sigma P = 670.416\), i.e. on average the rock mass produced 1m\(^{3}\) of seismic potency every 670.416 m\(^{3}\) of volume mined. The minimum single extraction was 19.5 m\(^{3}\), maximum 11116 m\(^{3}\), and the mean 1020.4 m\(^{3}\). The minimum distance between consecutive production blasts was 1 m, the maximum 355, and the mean 108 metres. The minimum time between production blasts was 0.96 days, the maximum 16, and the mean 1.63 days.
Figure 5.12 shows the cumulative number of events, cumulative potency, and cumulative volume mined vs. time for DataSet1. Size of the event scales with the radius of source volume with strain change \(\Delta \epsilon \geq 10^{-2}\), i.e. \(V=P/10^{-2}\), and the colour indicates the distance of that event from the \(\log P=2.24\) main shock. The size of the production blast scales with the size of the rectangle, and its colour scales with the distance of that blast from the main shock.
There is an increase in the rate of seismic events with \(\log P\geq -1.0\) from September 2007 to March 2008. While there is no increase in the rate of seismic activity before the \(\log P=1.82\) on 06 March 2009, there is an increase in the frequency of mid-size events which one could attribute to blasting of large volumes of rock. However, after January 2010, the mine also had a few large extractions, but there were only few mid-size events.
From January 2012, there was an increase in frequency of mid-size events, also reflected in the increased rate of CumP, that lasted almost to the \(\log P=2.24\) main shock on 13 June 2012. While the rate of production was steady over that time, the distances between subsequent extractions were smaller (Fig. 5.13 left) and at greater depth (Fig. 5.13 right).
Fig. 5.13
Distances between consecutive blasts (left) and elevation of production blasts vs. time (right) for DataSet1
Bild vergrößern
Seismic Activity and Production During DataSet2
This data set also starts on 07 September 2007 and ends on 07 July 2013 just after a \(\log P=2.61 \left (m=2.66\right )\) event. It spans 2130 days and includes 2818 events with \(\log P\geq -1.0 \left (m\geq 0.25\right )\), which gives the rate of 1.32 events/day. These events delivered \(\varSigma P = 2526.06\) m\(^{3}\) of seismic potency at the rate of 1.186 m\(^{3}\)/day, which is higher than that during DataSet1. The largest event in DataSet2 has \(\log P=2.61 \left (m=2.66\right )\), and there are 30 events with \(\log P\geq 1.0 \left (m\geq 1.59\right )\).
The mean recurrence interval \(\bar {t}\left (\log P\geq -1.0\right ) = 0.756\) days with the standard deviation of \(1.125\) days, which gives the coefficient of variation \(C_{v}\left (\log P\geq -1.0\right ) = 1.49\). The mean recurrence interval \(\bar {t}\left (\log P\geq 1.0\right ) = 73.44\) days with the standard deviation of \(94.76\) days, which gives the coefficient of variation \(C_{v}\left (\log P\geq 1.0\right ) = 1.29\).
During this time there were 1343 production blasts extracting 1346518 m\(^{3}\) of rock that gives the average extraction rate of 632.23 m\(^{3}\)/day. The ratio \(\Sigma V_{m}/\Sigma P = 533.06\), i.e. on average the rock mass produced 1m\(^{3}\) of seismic potency every 533.06 m\(^{3}\) of volume mined. The minimum single extraction was 19.5 m\(^{3}\), maximum 11116, and the mean 1002.6 m\(^{3}\). The minimum distance between consecutive production blasts was 1 m, the maximum 355, and the mean 105 metres. The minimum time between production blasts was 0.25 days, the maximum 16, and the mean 1.58 days.
Figure 5.14 shows the cumulative number of events, cumulative potency, and cumulative volume mined versus time for DataSet2. Here the vertical black line marks the end of DataSet1 that was analysed above. After the \(\log P=2.24\) on 13 June 2012, the rate seismic activity was fairly steady but characterised by frequent mid-size events that pushed the rate of cumulative potency up.
Fig. 5.14
Cumulative number of events (left), cumulative potency, and cumulative volume mined vs. time (right) for DataSet2
Bild vergrößern
Figure 5.15 shows the distances between consecutive events and the elevation of production blasts versus time for DataSet2.
Fig. 5.15
Distances between consecutive blasts (left) and elevation of production blasts vs. time (right) for DataSet2
Bild vergrößern
Comments
1.
The rate of rock extraction, measured by \(\varSigma V_{m}\)/day, was also similar, 628.1 m\(^{3}\)/day versus 633.1 m\(^{3}\)/day.
 
2.
The mean span between consecutive production blasts in DataSet1 was 368 m in X, 195 m in Y, and 245 in Z-direction.
 
3.
The mean span of 272 production blasts during the additional 389 days of mining past DataSet1 was 265 m in X, 154 m in Y, and 102 m in Z-direction.
 
4.
The volume mined weighted depth of production was 91 m deeper during the additional 389 days of mining.
 
5.
The rate of seismic activity, \(N\left (\log P\geq -1.0\right )\)/day, was similar, 1.31/day versus 1.32/day, but the rate \(N\left (\log P\geq 1.0\right )\)/day increased from 0.0115/day in DataSet1 to 0.0141/day in DataSet2.
 
One can postulate that apart from increasing extraction ratio, the more concentrated rock extraction at greater depth may have contributed to the observed increase in seismic hazard.
The Upper Limit to the Next Largest Event
The upper limit to the next record breaking potency, \(\log P_{max}\), was estimated from \(\log P_{max} = \log P_{maxo} + \Delta \log P_{max}\), where \(\log P_{maxo}\) is the maximum observed \(\log P\) and \(\Delta \log P_{max}\) is the maximum expected jump that can be estimated using order statistics based on the history of observed jumps in record \(\log P\),
$$\displaystyle \begin{aligned} \begin{array}{rcl} \Delta\log P_{max}& =&\displaystyle 2\max\left(\Delta\log P_{maxo}\right)\\ & &\displaystyle -\sum_{i=0}^{n-1}\left[\left(1-\frac{i}{n}\right)^{n}-\left(1-\frac{i+1}{n}\right)^{n}\right]\Delta\log P_{maxo-i}, \end{array} \end{aligned} $$
where \(\Delta \log P_{maxo-i}\) are the observed jumps in the history of records and n is the number of observed record jumps. Note that the order statistics estimate of \(P_{max}\) or \(\log P_{max}\) is independent of the underlying probability distribution, and therefore, one can make a reasonable estimate even if the data do not conform to the potency frequency power law. The upper limit to the next record breaking potency for DataSet1 \(\log P_{max}=2.68\) and for DataSet2 \(\log P_{max}=3.04\), see Fig. 5.17.
Size Distribution
Figure 5.16 shows the cumulative number of unbinned data vs. \(\log P\geq -1.0\) for DataSet1 (left) and DataSet2 (right). The black solid line represents the UT power law, \(N\left (\geq P\right ) = \alpha \left (P^{-\beta }-P_{max}^{-\beta }\right )\), assuming truncation at \(\log P_{max}\) and the red solid line at \(\log P_{nrb}\). The straight blue line represents the open-ended power law, \(N\left (\geq P\right ) = \alpha P^{-\beta }\), also known as the Gutenberg-Richter relation, and is plotted as a reference. Data points below \(\log P_{min}=-1.0\) are excluded from fitting. The grey dashed lines on both sides are the 95% confidence limits. The light grey vertical spikes below the data illustrate what would be the empirical pdf of the distribution for bin size 0.05.
Fig. 5.16
Potency frequency fits to DataSet1 (left) and DataSet2 (right) recorded at Mine D
Bild vergrößern
Fig. 5.17
History of the record breaking events and the estimated range of \( \log P_{nrb}\) for Mine D based on DataSet1 (left) and DataSet2 (right)
Bild vergrößern
Note that DataSet2 fits the UT power law a little better than DataSet1, although in both cases it overestimates frequency of events between \(\log P=1.2\) and \(\log P=2.0\).
The Expected Next Record Breaking Potency
The expected next record potency, \(P_{nrb}\), was calculated from
$$\displaystyle \begin{aligned} P_{nrb\left(k\right)}=\beta\left(P_{max}^{1-\beta}-P_{r\left(k-1\right)}^{1-\beta}\right)/\left[\left(1-\beta\right)\left(P_{r\left(k-1\right)}^{-\beta}-P_{max}^{-\beta}\right)\right], \end{aligned}$$
where \(P_{r\left (k-1\right )}\) is the potency of the previous (or the last) record breaking event and \(\beta \) is the exponent of the potency frequency distribution.
Figure 5.17 (left) shows the history of records and the estimated range of the next record \(\log P\) for DataSet1 with colour indicating the distance to the main shock. The record breaking event \(\log P=2.61\) recorded on 07 July 2013 is within the predicted range of \(2.45 \leq \log P_{nrb} \leq 2.68\). There are eight forward counting \(\log P\) records and one backward counting \(\log P\) record, which indicates an upward trend in seismic hazard. Figure 5.17 (right) shows the history of records and the estimated range of the next record breaking \(\log P\) for DataSet2. The expected range of the next record breaking event is \(2.81 \leq \log P_{nrb} \leq 3.04\). There are nine forward counting \(\log P\) records and one backward counting record, confirming an upward trend in seismic hazard.
Intermediate Term Hazard Change in Time
The intermediate term hazard, \(\Pr \left [\geq \log P;\Delta T\right ]\), was calculated assuming the stationary Poisson process, \(\Pr \left (\geq P,\Delta T\right ) = 1 - \exp \left [-\left (\Delta T/\Delta t\right )N\left (\geq P\right )\right ]\), where \(\Delta t\) is the time of observation, \(N\left (\geq P\right ) = \alpha \left (P^{-\beta }-P_{Limit}^{-\beta }\right )\), \(P_{Limit}\) is \(\log P_{nrb}\) for the lower limit (red line) and \(\log P_{max}\) for the upper limit (black line) of probabilities, and \(\alpha \) and \(\beta \) are the respective parameters of the upper truncated potency frequency distribution.
Figure 5.18 shows the intermediate term probabilities of having at least one event greater than or equal to a given \(\log P\) over 3, 6, and 12 months.
Fig. 5.18
Intermediate term probabilities for DataSet1 (left) and DataSet2 (right)
Bild vergrößern
Comments
1.
Seismic hazard for the period 07 September 2007 to 07 July 2013 (DataSet2, \(\Delta t = 2130\) day) is higher than it was during the period 07 September 2007 to 13 June 2012 (DataSet1, \(\Delta t = 1741\) days).
 
2.
The ratio \(\varSigma V_{m}/\varSigma P\) indicates how much rock needs to be extracted to produce 1 m\(^{3}\) of seismic potency. During the first 1741 days of mining, this ratio was 670.42, and it dropped to 245.1 during the additional 389 days of mining.
 
3.
During the first 1741 days, the activity rate of events with \(\log P\geq 1.0\) was 0.0115/day, and it increased to 0.0283/day during the additional 389 days.
 
4.
During the first 1741 days, the probability of having at least one event with \(\log P\geq 2.0 \left (m\geq 2.25\right )\) within 1 year was 0.407, and it increased to 0.475 during the additional 389 days of mining.
 
Frequently mine management asks for seismic hazard to be presented as seismic hazard curve that shows the annual rate at which a given \(\log P\) will be exceeded. Hazard curve is then superimposed on the probability rating scheme, or risk matrix, defined by the mine. Figure 5.19 shows hazard curves, i.e. the expected annual rate of exceedance for both data sets.
Fig. 5.19
Hazard curves for DataSet1 (left) and DataSet2 (right)
Bild vergrößern

5.6 Seismic Response to Step Loading: Short Term Hazard

5.6.1 Introduction

The rock mass dynamics in mines is driven mainly by the transient deformation of nearby excavations in response to extraction of stressed rock and by the near-field deformation of seismic sources, including pillar and abutment failures, and by blasting. To a lesser degree, it is also influenced by waves from remote sources of seismic radiation and by tidal forces. The step loading caused by rock extraction and by seismic events induces large displacements, plastic instability, splitting, buckling, and bursting of rock close to excavations. The energy imparted into the rock mass during step loading is partially converted into strain energy and partly dissipated through friction, local plastic deformation, and strain energy of radiated waves.
After a step loading, one can observe two partly competing stress-related processes occurring in the affected rock mass:
(1) An excitation phase that elevates the average stress level in volumes of rock where loading is faster than the ability to dissipate the excess of strain energy. It is generally short-lived, and it can be observed by monitoring changes in seismically inferred stress and strain by high-resolution networks.
(2) A relaxation phase that cascades the system down from its excited state through intermediate phases toward a non-equilibrium steady state. It is facilitated by different forms of inelastic deformation—seismic and aseismic. It modifies the stress pattern by reducing its elevated levels and moving it further away from excavations. It generates the bulk of the seismic activity after loading, and it lasts much longer than the excitation phase.
Seismic rock mass response to blasting is driven strongly by the stress level in rock surrounding the blast and by the volume of rock blasted. In the case of a preconditioning blast, when there is little or no rock extraction, the stress level plays the major role. In most cases, rock extraction by blasting induces and triggers seismic events immediately and in close proximity to the blast, which helps to set up a proper re-entry protocol, i.e. the exclusion zone and re-entry time.
Aftershock sequences after larger seismic events in mines or after major production blasts are not as well developed and do not last as long as after tectonic earthquakes of similar size. The main reason is faster relaxation facilitated by the presence of openings and extensive fracturing. If the immediate aftershocks are subdued and locate close to the main shock, there is less potential for a large event. As the immediate aftershocks extend further away from the main shock, it is more likely they may trigger larger event, see Fig. 5.20.
Fig. 5.20
Cumulative number (left) and distances (right) of seismic events to the main shock (MS) with \( \log P=2.61\)
Bild vergrößern
However, rock mass convergence associated with large rock extractions or with larger pillar removals may produce elevated seismic response for days or even weeks with sporadic elevated levels of seismic activity away from the place of extraction. Locations and orientations of larger events during aftershock sequences are controlled by geological structures rather than by the elevated stress level.
The seismic activity after step loading tends to cluster in space and time. In some cases, the rate of activity follows typical aftershock sequences that can be described by the Omori law or the stretched exponential relaxation function. However, in some cases, the response is more complex and cannot be described by a simple intensity function and therefore needs to be treated with non-parametric statistics.

5.6.2 Seismicity Rate Change

The fundamental premise here is that as the rate of seismic activity increases so does the likelihood that one of these events may be larger and damaging. In many cases, specifically after production or development blasts, the recorded activity is complex and does not follow the typical aftershock sequences that could be described by the Omori law or the stretched exponential relaxation function, and therefore needs to be treated by non-parametric statistics. One way to check if the elevated seismic activity has returned to an acceptable level is to test the null hypothesis of no change. The acceptable level of seismic activity can be defined a priori, or it can be taken as an average activity rate before step loading.
To detect changes in seismic activity rates in two different time intervals, \(\Delta t_{1}\) and \(\Delta t_{2}\), within the same volume of rock, one can count the respective number of recorded events above a certain potency. If the time intervals are equal and relatively long and the observed number of events is significantly different, then a statement can be made about the relative change. However, the associated uncertainty increases as the time intervals get shorter and as the difference in the event counts becomes smaller. The situation is even more difficult if the time intervals are not equal.
Under normal conditions, the event counts can be considered as outcomes of a Poisson process, and therefore, their occurrence can be very irregular. To measure the seismicity rate change, we need the probability density function of the activity rate, \(\lambda \). This can be derived by normalising the Poissonian density function (see Eq. 5.5) so that the integral over \(\lambda \) from 0 to \(\infty \) is unity, which gives
$$\displaystyle \begin{aligned} f\left(\lambda\right)=\Delta t\left(\lambda\Delta t\right)^{N}\exp\left(-\lambda\Delta t\right)/N!{} \end{aligned} $$
(5.16)
The probability that the seismicity rate in two different time intervals, with densities \(f_{1}\) corresponding to time interval \(\Delta t_{1}\) and \(f_{2}\) to \(\Delta t_{2}\), increased by more than k times is \(\Pr \left [\left (\lambda _{2}/\lambda _{1}\right )>k\right ] = \intop _{0}^{\infty } d\lambda _{1} f_{1}\left (\lambda _{1}\right ) \intop _{k\lambda _{1}}^{\infty } d\lambda _{2} f_{2}\left (\lambda _{2}\right )\), which, taking \(\lambda _{1} = N_{1}\left (\geq \log P\right )/\Delta t_{1}\) and \(\lambda _{2} = N_{2}\left (\geq \log P\right )/\Delta t_{2}\), gives
$$\displaystyle \begin{aligned} \Pr\left(\frac{\lambda_{2}}{\lambda_{1}}>k\right)=\frac{1}{N_{2}!N_{1}!}\intop_{0}^{\infty}x^{N_{1}}\exp\left(-x\right)\varGamma\left(N_{2}+1,kx\frac{\Delta t_{2}}{\Delta t_{1}}\right)dx,{} \end{aligned} $$
(5.17)
where \(\varGamma \left (N_{2}+1,kx\Delta t_{2}/\Delta t_{1}\right ) = \intop _{kx\Delta t_{2}/\Delta t_{1}}^{\infty }\exp \left (-t\right )t^{N_{2}+1}dt\) is the upper incomplete Gamma function, and the ratio \(\Delta t_{2}/\Delta t_{1}\) caters for the unequal time intervals (Marsan, 2003).
Illustrative Example
Let us assume that in a given volume of rock, \(\Delta V\), during \(\Delta t_{1}=10\) days preceding the step loading in a remote section of a mine, the seismic system recorded \(N_{1}=10\) events with \(\log P\geq -1.0\), i.e. \(\lambda _{1}=1.0\). Figure 5.21 (left) shows the probability calculated by Eq. (5.17) that the event rate increased by more than k times, if during \(\Delta t_{2}=10\) days after blasting the system recorded \(N_{2}=10\), 15 or 20 events with \(\log P\geq -1.0\), i.e. \(\lambda _{2}= 1\), \(1.5\), or \(2.0\), respectively. The \(\Pr \left [\left (\lambda _{2}/\lambda _{1}\right )>k=1\right ] = 0.5\) indicates an equal chance of activation (\(k > 1\)) or slow down (\(k < 1\)) of seismic activity in \(\Delta V\). This is the case when \(\Delta t_{2} = \Delta t_{1}\) and \(\lambda _{1} = \lambda _{2}\), see the green dot in Fig. 5.21 (left).
Fig. 5.21
Probabilities of an increase in seismicity rate for equal (left) and unequal (right) time intervals
Bild vergrößern
If, however, it had taken \(\Delta t_{1}=20\) days to record \(N_{1} = 10\) events with \(\log P\geq -1.0\) before the step loading, then with all other parameters being equal, the probability of activation would increase to \(\Pr \left [\left (\lambda _{2}/\lambda _{1}\right )>k=1\right ] = 0.944\), the probability that seismic activity increased by more than \(k = 1.5\) times would be \(\Pr \left [\left (\lambda _{2}/\lambda _{1}\right )>k=1.5\right ] = 0.747\), and the probability that seismic activity at least doubled \(\Pr \left [\left (\lambda _{2}/\lambda _{1}\right )>k=2\right ] = 0.5\), see the green dots at Fig. 5.21 (right). Then, the probability that \(0.944\) or \(0.747\) or more could be obtained by chance if there was no change, which is the null hypothesis, is \(1 - 0.944 = 0.056\) or \(1 - 0.744 = 0.256\), respectively.
One can also solve Eq. (5.17) for k for a given probability, e.g.
$$\displaystyle \begin{aligned} \left(N_{1}!N_{2}!\right)^{-1}\intop_{0}^{\infty}x^{N_{1}}\exp\left(-x\right)\varGamma\left[N_{2}+1,kx\left(\Delta t_{2}/\Delta t_{1}\right)\right]dx=0.9,{} \end{aligned} $$
(5.18)
to answer the following question: What is the ratio of seismicity rate change that gives 90% certainty. For example, if during \(\Delta t_{1}=10\) days preceding the step loading in a remote section of a mine, the seismic system recorded \(N_{1}=10\) events with \(\log P\geq -1.0\), and \(N_{2} = 20\) events during \(\Delta t_{2}=10\) days after, then there is 90% probability that seismicity rate changed by at least \(k = 1.2\) times.
Reference Activity Level
Equation (5.17) can be applied to monitor if the elevated rate of seismic activity after production blasts returned to an acceptable reference level. Note that frequently activity rate before larger events is elevated and, if used as a reference, would underestimate the re-entry time. The reference activity rate can be estimated by taking an average over periods of times that satisfy the following criteria: (a) they are outside the influence of blasting, (b) there were no larger events, and (c) there was normal production activity and people working in the area. It is expected that the coefficient of variation of the data selected to estimate the reference activity will not be far from 1.0, so the reference activity would not deviate too far from Poissonian.
Typically, the exclusion zone and time after a large seismic event, or blast, have the following characteristics. (1) The exclusion time increases as the size of the main shock increases. (2) The exclusion zone increases as the size of the main shock increases. (3) The exclusion time decreases as the distance to the main shock increases.
Data
One can use the following types of data to measure seismic activity:
1.
Magnitude, or \(\log P\) or \(\log E\), of associated seismic events. This data is delayed since it requires seismological processing, i.e. location and source parameters. Moreover, to provide a reasonable location, the seismic system accepts events that associated with at least five stations, and this removes a great number of small events from the analysis. Data on the activity of associated events also underreports on immediate aftershocks, some of them buried in the coda of the main shock.
 
2.
The peak ground velocity, PGV , and/or the cumulative absolute displacement, CAD, which is the integral of the absolute value of a velocity time series, \(CAD = \intop _{0}^{t}|\mbox{v}\left (t\right )|dt\), and has units of displacement. Both parameters can be extracted from a stream of continuous data provided by seismic system. This data is available in real time since it does not require processing. An event is declared when during a given time interval, say \(\Delta t=0.25\) second, the GM parameter exceeds a predefined threshold.
 
The data for analysis can be derived from a predefined polygon which is defined as a seismogenic volume that generates seismicity affecting working places. Therefore, all relevant parameters are derived from the data selected from this polygon. This method has been widely applied in mines for many years. The most difficult, and also subjective task, here is the definition of the polygon. Different polygons select different data sets and therefore will produce different seismic characteristics.
Alternatively, one can apply the polygon-less or the influence based approach where one takes into account the influence of all available seismic events, regardless of their location, on a particular working place. The preferred measures of influence can be PGV  and/or CAD since their influence is moderated by the distance from the seismic source to the place of potential exposure.

5.6.3 Omori Relaxation Function

In 1893 Fusakichi Omori published the 11 page long note “On aftershocks” (Omori, 1893), which was the abstract of the paper published a year later (Omori, 1894), where he described the activity of aftershocks at any time x after the main shock as \(y = k\left (x+h\right )^{-1}\), where k and h are constant, which “represents a rectangular hyperbola and implies that the activity varies nearly inversely with time”. Hirano (1924) and Utsu (1957) modified Omori’s equation by introducing the exponent p, and today the Omori law is presented as
$$\displaystyle \begin{aligned} dN/dt=k\left(t+c\right)^{-p},{} \end{aligned} $$
(5.19)
where \(k, c\), and p are parameters and t is time. The total number of events \(N_{t} = N\left (0,\infty \right )\) is derived from \(N_{t}=\int _{0}^{\infty }k\left (t+c\right )^{-p}dt\) that gives \(N_{t} = k c^{1-p} / \left (p-1\right )\) for \(p>1\) and \(N_{t} = \infty \) for \(p \leq 1\) (Fig. 5.22).
Fig. 5.22
Omori relaxation for \(k=100\), \(c=0.1\) and different values of parameter p on the log-log scale
Bild vergrößern
The parameter k depends on the magnitude of the main shock and for given c and p scales with the total number of events. The parameter c is the offset time that accounts for the incompleteness of the data set immediately after the main shock—the better the resolution of the seismic network, the smaller the c. Setting \(c > 0\) also prevents the infinite N at \(t=0\), and its physical meaning, however, is debatable.
When plotting the observed number of events within a certain interval of time versus time on a log-log scale, the data points should tend to a straight line whose slope is an estimate of p. The parameter p controls the rate of decay of aftershocks, and, for slowly decaying sequences with \(p\leq 1\), the Omori law predicts that the total number of events becomes unbounded with increasing time. The parameter p measures the overall relaxation time of the process following the major event. The lower the p-value, the slower the relaxation, which is characteristic of stiffer systems; the opposite would apply to softer systems. As a consequence, the p-value would be expected to correlate negatively with the Gutenberg-Richter b-value, e.g. (Wang, 1994). Note that p and c are positively and almost linearly correlated, which can frustrate the estimation process. In general, the Omori law is an open-ended power law, \(N\left (\geq t\right ) \sim t^{-p}\), has infinite moments and therefore the property of scale invariance, i.e. the relative change of \(N\left (\geq st\right ) / N\left (\geq t\right ) = s^{-p}\) is independent of t, and therefore it lacks a characteristic scale—there is no characteristic time. To be physical, the power law description of a finite system must be bounded; otherwise, aftershocks would continue forever.
The number of events in a time interval \(N\left (t_{1},t_{2}\right )\) after the step loading can be calculated by \(N(t_{1},t_{2})=\int _{t_{1}}^{t_{2}}k/\left (t+c\right )^{p}dt\), which gives
$$\displaystyle \begin{aligned} N(t_{1},t_{2})=\left\{ \begin{array}{cc} \frac{k}{(1-p)}\left[\left(t_{2}+c\right)^{1-p}-\left(t_{1}+c\right)^{1-p}\right], & p\neq1\\ k\ln\left(\frac{t_{2}+c}{t_{1}+c}\right), & p=1 \end{array}\right.{} \end{aligned} $$
(5.20)
and the ratio
$$\displaystyle \begin{aligned} \frac{N\left(t_{1},t_{2}\right)}{N\left(0,t_{2}\right)}=\left\{ \begin{array}{cc} \left[\left(t_{2}+c\right)^{1-p}-\left(t_{1}+c\right)^{1-p}\right]/\left[\left(t_{2}+c\right)^{1-p}-c^{1-p}\right], & p\neq1\\ \ln\left[\left(t_{2}+c\right)/\left(t_{1}+c\right)\right]/\ln\left[\left(t_{2}+c\right)/c\right], & p=1 \end{array}\right..{} \end{aligned} $$
(5.21)

5.6.4 Omori Distribution, Parameters, and Simple Omori

To get the probability distribution, we need to normalise parameter k so \(\int _{0}^{T}k\left (t+c\right )^{-p}dt =1\), where T is the assumed end of the aftershock sequence, which gives
$$\displaystyle \begin{aligned} \left\{ \begin{array}{cc} f\left(t\right)=\left(p-1\right)\left(t+c\right)^{-p}/\left[c^{1-p}-\left(T+c\right)^{1-p}\right], & p\neq1\\ f\left(t\right)=\left(t+c\right)^{-1}/\ln\left[T/\left(1+c\right)\right], & p=1 \end{array}\right.,{} \end{aligned} $$
(5.22)
and the cumulative distribution function, \(F\left (t\right ) = \Pr \left (\leq t\right ) = \intop _{0}^{t}f\left (u\right )du\), which gives
$$\displaystyle \begin{aligned} \left\{\!\begin{array}{cc} \Pr\left(\leq t\right)\!=\!c^{-p}\left(T\!+\!c\right)^{-p}\left[c\left(T\!+\!c\right)\!-\!c^{p}\left(T\!+\!c\right)\right]/\left[c^{1-p}\!-\!\left(T\!+\!c\right)^{1-p}\right], & p\neq1\\ \Pr\left(\leq t\right)=\ln\left[\left(c+t\right)/c\right]/\ln\left[t/\left(1+c\right)\right], & p=1 \end{array}\right.,{} \end{aligned} $$
(5.23)
and \(\Pr \left (\leq t\right )=\left (t/T\right )^{1-p}\) for \(c = 0\) and \(p \neq 1\). The probability of having an event during the time interval \(\left (t_{1},t_{2}\right )\) is
$$\displaystyle \begin{aligned} \left\{ \begin{array}{cc} \Pr\left(t_{1},t_{2}\right)=\left[\left(t_{1}+c\right)^{1-p}-\left(t_{2}+c\right)^{1-p}\right]/\left[c^{1-p}-\left(T+c\right)^{1-p}\right], & p\neq1\\ \Pr\left(t_{1},t_{2}\right)=\ln\left[\left(t_{2}+c\right)/\left(t_{1}+c\right)\right]/\ln\left[T/\left(1+c\right)\right], & p=1 \end{array}\right..{} \end{aligned} $$
(5.24)
For \(p \leq 1\), the end of the aftershock sequence T must be finite; otherwise, the integral diverges, but for \(p > 1\) there is a finite number of aftershocks as \(T \rightarrow \infty \) and the respective equations are as follows:
$$\displaystyle \begin{aligned} \left\{ \begin{array}{cc} f\left(t\right)=\left(p-1\right)\left(t+c\right)^{-p}/c^{1-p}, & p>1\\ F\left(t\right)=-\left[\left(t-c\right)^{1-p}-c^{1-p}\right]/c^{1-p}, & p>1\\ \Pr\left(t_{1},t_{2}\right)=\left[\left(t_{1}+c\right)^{1-p}-\left(t_{2}+c\right)^{1-p}\right]/c^{1-p}, & p>1 \end{array}\right..{} \end{aligned} $$
(5.25)
The maximum likelihood (ML) method of estimating parameters of the Omori relation is similar to the method of calculating the \(\beta \)-value of the power law size distribution. However, the results are not always stable, because p and c are almost linearly correlated. One can simplify calculations by assuming that parameter \(c=0\), i.e. simple Omori (SO). Then to get the probability density function, we need to normalise parameter k so \(\intop _{t_{1}}^{t_{2}}kt^{-p} = 1\), which gives \(k = \left (1-p\right ) / \left (t_{2}^{1-p}-t_{1}^{1-p}\right )\). The ML function then is \(L\left (t,p\right ) = \prod _{j=1}^{N}\left (1-p\right ) t_{j}^{-p} / \left (t_{2}^{1-p}-t_{1}^{1-p}\right )\), where N is the number of events in the time period \(t_{2} - t_{1}\). The parameter p can be calculated from \(\partial \ln L\left (p\right )/\partial p = 0\), which gives
$$\displaystyle \begin{aligned} \frac{N}{\left(1-p\right)}+\frac{N\left(t_{2}^{1-p}\ln t_{2}-t_{1}^{1-p}\ln t_{1}\right)}{t_{1}^{1-p}-t_{2}^{1-p}}+\sum_{j=1}^{N}\ln t_{dj}=0,{} \end{aligned} $$
(5.26)
where \(t_{dj}\) is the time of seismic events. Equation (5.26) needs to be solved numerically for p. The parameter k can then be calculated from
$$\displaystyle \begin{aligned} N=k\intop_{t_{1}}^{t_{2}}t^{-p}dt=\frac{k}{(1-p)}\left[t_{2}^{1-p}-t_{1}^{1-p}\right]\quad \Rightarrow\quad k=\frac{N\left(1-p\right)}{t_{2}^{1-p}-t_{1}^{1-p}}.{} \end{aligned} $$
(5.27)
The standard deviation of p can be obtained from the second derivative of the log likelihood function, \(\left [-\partial ^{2}\ln L\left (p\right )/\partial p^{2}\right ]^{-1/2}\), which gives
$$\displaystyle \begin{aligned} s_{d}\left(p\right)=\left\{ N/\left(1-p\right)^{2}-N\left(t_{2}/t_{1}\right)^{1-p}\left[\ln\left(t_{2}/t_{1}\right)\right]^{2}/\left[\left(t_{2}/t_{1}\right)^{1-p}-1\right]^{2}\right\} ^{-1/2}.{} \end{aligned} $$
(5.28)
The single most important factor influencing estimates of \(s_{d}(p)\) is the number of observations, and for small N, the estimates are unreliable. For \(p=1\), taking \(\partial \ln L(p=1,k)/\partial k = 0\) gives \(k = N / \left (\ln t_{2}\ln t_{1}\right )\).

5.6.5 Stretched Exponential Relaxation Function

An alternative to the Omori law is the stretched exponential function introduced by von Rudolf Kohlrausch (1854) to interpret charge relaxation in a Leiden jar, which was the first capacitor. The stretched exponential is also known as the Kohlrausch-Williams-Watts (KWW) function, after Williams and Watts (1970) used it to characterise the dielectric relaxation rates in polymers. The stretched exponential function has a fatter tail than the exponential, but less so than a pure power law distribution, offering a compromise between the two descriptions. It can therefore be used to account for both a limited scaling regime and a cross over to non-scaling. While the Omori law is purely empirical, the parameters of the stretched exponential have a clear physical meaning, i.e. the relaxation time of the exponential decay process and the total number of events.
The function is defined as \(\exp \left [-\left (t/\tau \right )^{q}\right ]\) for \(t \geq 0\), where \(\tau > 0\) is the relaxation time and \(0<q<1\) is called the stretching or the shape exponent. The origin of the stretched exponential is assumed to be the result of a competition between two or more exponential processes. The reciprocal of q is called a heterogeneity parameter, \(h = 1/q\), which is the number of generations in a multiplicative process, and it gives an insight into the heterogeneity in complex systems. The larger the values of relaxation time \(\tau \), the slower the relaxation processes. The smaller the value of the shape exponent \(q < 1\), the quicker the decay for short times and slower for long times.
To apply the stretched exponential to aftershock sequences, it is convenient to start at time \(t_{1}\) after \(t_{0}\) when the step loading was applied (Kisslinger, 1993). Then the number of events yet to occur in a sequence at time \(t_{1}\) is
$$\displaystyle \begin{aligned} N\left(t_{1},T\right)=N\left(t_{0},T\right)\exp\left[-\left(t_{1}/\tau\right)^{q}\right],{} \end{aligned} $$
(5.29)
where \(N\left (t_{0},T\right )\) is the total number of events that will eventually occur and \(T - t_{0}\) is the duration of the process (Fig. 5.23). At time \(t_{1} = \tau \), the number of events still to occur is \(N\left (\tau ,T\right ) = N\left (t_{0},T\right )/e\), regardless of q, so 63% of the work is done during the relaxation time. The number of events that had occurred to time \(t_{1}\) is \(N(t_{0},t_{1}) = N\left (t_{0},T\right ) - N(t_{1},T)\), which gives a prediction of the total number of events at time \(t_{1}\) as
$$\displaystyle \begin{aligned} N\left(t_{0},T\right)=N\left(t_{0},t_{1}\right)/\left\{ 1-\exp\left[-\left(t_{1}/\tau\right)^{q}\right]\right\} .{} \end{aligned} $$
(5.30)
Fig. 5.23
Cumulative number of events vs. time after step loading for \(\tau =4\) hours and for selected q(left), and for \(q=0.5\) and for selected \(\tau \)(right)
Bild vergrößern
The expected number of events to occur in the interval \(\left (t_{1},t_{1}+\Delta T\right )\) is
$$\displaystyle \begin{aligned} N\left(t_{1},t_{1}+\Delta T\right)=N\left(t_{0},T\right)\left\{ \exp\left[-\left(t_{1}/\tau\right)^{q}\right]-\exp\left[-\left(\left(t_{1}+\Delta T\right)/\tau\right)^{q}\right]\right\} .{} \end{aligned} $$
(5.31)

5.6.6 Stretched Exponential Distribution and Its Parameters

The cumulative distribution function is
$$\displaystyle \begin{aligned} F\left(t\right)=\Pr\left(\leq t\right)=N\left(t_{0},t\right)/N\left(t_{0},T\right)=1-\exp\left[-\left(t/\tau\right)^{q}\right],{} \end{aligned} $$
(5.32)
and the survival function, \(S\left (t\right ) = \exp \left [-\left (t/\tau \right )^{q}\right ]\). The probability density function is
$$\displaystyle \begin{aligned} f\left(t\right)=dF\left(t\right)/dt=\left(q/\tau\right)\left(t/\tau\right)^{q-1}\exp\left[-\left(t/\tau\right)^{q}\right].{} \end{aligned} $$
(5.33)
The probability of having an event in the time interval \(\left (t,t+\Delta T\right )\) is \(\Pr \left (t,t+\Delta T\right ) = F\left (t+\Delta T\right ) - F\left (t\right ) = \exp \left [-\left (t/\tau \right )^{q}\right ] - \exp \left [-\left (\left (t+\Delta T\right )/\tau \right )^{q}\right ]\), and the expected number of events in this time interval is \(N\left (t,t+\Delta T\right ) = N\left (t_{0},T\right ) \Pr \left (t,t+\Delta T\right )\). For \(q = 1\), the stretched exponential becomes a simple exponential distribution and for \(q < 1\) decays slower than exponential for large times (thick tail). For \(q > 1\), it decays faster than exponential and is known as the Weibull distribution. For \(q = 2\), it is also known as the Rayleigh distribution.
All moments of the stretched exponential distribution are finite, the mean, or the expected value is \(\tau \varGamma \left (1+1/q\right ) = \left (\tau /q\right ) \varGamma \left (1/q\right )\), and variance \(\tau ^{2} \varGamma \left (1+2/q\right ) - \tau ^{2} \varGamma \left (1+1/q\right )^{2}\). The gamma function, \(\varGamma \left (t\right ) = \intop _{0}^{\infty }x^{t-1}e^{-x}dx\), reduces to \(\varGamma \left (n\right ) = \left (n-1\right )!\) for positive integers and \(\varGamma \left (t+1\right ) = t\varGamma \left (t\right )\). The median is \(F\left (t\right ) = \exp \left [-\left (t_{0.5}/\tau \right )^{q}\right ] = 0.5\), which gives \(t_{0.5} = \tau \left (\ln 2\right )^{1/q}\).
The function \(t\cdot f\left (t\right ) = q \left (t/\tau \right )^{q} \exp \left [-\left (t/\tau \right )^{q}\right ]\) has a maximum at the relaxation time \(\tau \), regardless of q, and the \(\tau \cdot f\left (\tau \right ) = q/e\) depends only on q, see Fig. 5.24. These two properties of \(t\cdot f\left (t\right )\) may be utilised while fitting the data.
Fig. 5.24
\(t\cdot f \left (t \right )\) for \(\tau =4\) hours and for selected q(left), and for \(q=0.5\) and for selected \(\tau \)(right)
Bild vergrößern
The maximum likelihood function is
$$\displaystyle \begin{aligned} L\left(t_{j};\tau,q\right)=\prod_{j=1}^{n}f\left(t_{j}\right)=\prod_{j=1}^{n}\left(q/\tau\right)\left(t_{j}/\tau\right)^{q-1}\exp\left[-\left(t_{j}/\tau\right)^{q}\right], \end{aligned}$$
and \(\ln L\left (t_{j};\tau ,q\right ) = \sum _{j=1}^{n} \ln f\left (t_{j}\right ) = n\ln \left (q\right ) + q\sum _{j=1}^{n}\ln \left (t_{j}\right ) - \sum _{j=1}^{n}\ln \left (t_{j}\right ) - n q \ln \left (\tau \right ) - \sum _{j=1}^{n}\left (t_{j}/\tau \right )^{q}\). Taking \(\partial \ln L\left (t_{j};\tau ,q\right ) / \partial \tau = 0\) and \(\partial \ln L\left (t_{j};\tau ,q\right ) / \)\(\partial q = 0\) gives
$$\displaystyle \begin{aligned} \tau=\left[\left(1/n\right)\sum\left(t_{j}\right)^{q}\right]^{1/q},{} \end{aligned} $$
(5.34)
$$\displaystyle \begin{aligned} \frac{n}{q}-n\ln\left(\tau\right)+\sum\ln\left(t_{j}\right)-\tau^{-q}\left\{ \sum\left[\left(t_{j}\right)^{q}\ln\left(t_{j}\right)\right]-\ln\left(\tau\right)\sum\left(t_{j}\right)^{q}\right\} =0,{} \end{aligned} $$
(5.35)
respectively, where all sums are \(\sum _{j=1}^{n}\). Inserting \(\tau \) given by Eq. (5.34) into Eq. (5.35) gives
$$\displaystyle \begin{aligned} \left(1/q\right)+\left(1/n\right)\sum\ln\left(t_{j}\right)-\sum\left[\left(t_{j}\right)^{q}\ln\left(t_{j}\right)\right]/\sum\left(t_{j}\right)^{q}=0,{} \end{aligned} $$
(5.36)
which depends only on q but does not have an analytical solution and needs to be solved numerically, and then one can solve for \(\tau \) using Eq. (5.34). The standard deviation of q and \(\tau \) can be obtained from the second derivative of the log likelihood function, \(\left [-\partial ^{2}\ln L\left (q,\tau \right )/\partial q^{2}\right ]^{-1/2}\) and \(\left [-\partial ^{2}\ln L\left (q,\tau \right )/\partial \tau ^{2}\right ]^{-1/2}\), which gives
$$\displaystyle \begin{aligned} \begin{array}{rcl} s_{d}\left(q\right)=\left\{ \frac{n}{q^{2}}+\frac{1}{\tau^{q}}\sum t_{j}^{q}\left[\ln\left(t_{j}/\tau\right)\right]^{2}\right\} ^{-1/2}\ \text{and}\\ s_{d}\left(\tau\right)=\tau\left[q\left(n+\frac{q+1}{\tau^{q}}\sum t_{j}^{q}\right)\right]^{-1/2}.{} \end{array} \end{aligned} $$
(5.37)

5.6.7 Hazard Function, Conditional Probability, and Stretched Exponential

The hazard function is defined as \(h\left (t_{1}\right ) = f\left (t_{1}\right ) / \left [1-F\left (t_{1}\right )\right ] = f\left (t_{1}\right ) / S\left (t_{1}\right )\), where \(f\left (t_{1}\right )\) is the probability density function, \(F\left (t_{1}\right ) = \Pr \left (\leq t_{1}\right )\) is the cumulative distribution function, i.e. the probability that the event will happen before time \(t_{1}\), and \(S\left (t_{1}\right ) = \Pr \left (\geq t_{1}\right ) = 1 - F\left (t_{1}\right )\) is the survival function, i.e. the probability that the event will not happen until time \(t_{1}\).
If the distribution of the recurrence times is a stretched exponential with \(f\left (t_{1}\right ) = \left (q/\tau \right ) \left (t_{1}/\tau \right )^{q-1} \exp \left [-\left (t_{1}/\tau \right )^{q}\right ]\) and the cumulative distribution function \(F\left (t_{1}\right ) = 1-\exp \left [-\left (t_{1}/\tau \right )^{q}\right ]\), then the hazard function becomes
$$\displaystyle \begin{aligned} h\left(t_{1}\right)=\left(q/\tau\right)\left(t_{1}/\tau\right)^{q-1},{} \end{aligned} $$
(5.38)
where \(t_{1}\) is the time since the last event. For systems characterised by \(q=1\), the probability of having another event of similar size is independent of time, i.e. the stretched exponential distribution is reduced to the exponential one, and since the system has no memory, events occur randomly, see Fig. 5.25 (left).
Fig. 5.25
Hazard function (left) and conditional probability of the stretched exponential distribution (right)
Bild vergrößern
For \(q>1\), the probability of having another event of that size is increasing as the time since the last event increases. Interestingly, for \(q<1\), the probability of having another event of that size decreases with an increasing time since the last event.
For the intermediate and the long term hazard, one needs to consider the conditional probability that such an event will occur during \(\Delta T\), if the time since the last event of that size is \(t_{1}\), \(\Pr (\geq P;t_{1}\leq t\leq t_{1}+\Delta T\mid t_{1}<t) = \int _{t_{1}}^{t_{1}+\Delta T}f(t)dt/\int _{t_{1}}^{\infty }f(t)dt\), which, assuming the stretched exponential distribution, gives
$$\displaystyle \begin{aligned} \begin{array}{rcl} \Pr(\geq P;t_{1}\leq t\leq t_{1}+\Delta T\mid t_{1}<t)=F\left(\Delta T|t_{1}\right)\\ =1-\exp\left[\left(\frac{t_{1}}{\tau}\right)^{q}-\left(\frac{t_{1}+\Delta T}{\tau}\right)^{q}\right],{} \end{array} \end{aligned} $$
(5.39)
see Fig. 5.25 (right). The waiting time corresponding to a probability of 0.5 gives the median waiting time,
$$\displaystyle \begin{aligned} \Pr(\geq P;t_{1}\leq t\leq t_{1}+\Delta T\mid t_{1}<t)=0.5\;\Rightarrow\;\Delta T_{0.5}=\tau\left[\left(\frac{t_{1}}{\tau}\right)^{q}-\ln2\right]^{1/q}-t_{1}.{} \end{aligned} $$
(5.40)
The respective probability density function is the derivative \(d\left [F\left (\Delta T|t_{1}\right )\right ]d\left (\Delta T\right )\), which gives \(f\left (\Delta T|t_{1}\right ) = f\left (t_{1}+\Delta T\right ) / S\left (t_{1}\right )\), where \(S\left (t_{1}\right ) = 1 - F\left (t_{1}\right )\), and for the stretched exponential,
$$\displaystyle \begin{aligned} f\left(\Delta T|t_{1}\right)=\frac{q}{\tau}\left(\frac{t_{1}+\Delta T}{\tau}\right)^{q-1}\exp\left[\left(\frac{t_{1}}{\tau}\right)^{q}-\left(\frac{t_{1}+\Delta T}{\tau}\right)^{q}\right].{} \end{aligned} $$
(5.41)
The expected time to the next event then is \(\left \langle \Delta T\right \rangle = \left [1/S\left (t_{1}\right )\right ] \intop _{0}^{\infty }\Delta Tf\left (t_{1}+\Delta T\right )d\)\(\left (\Delta T\right )\), which for the stretched exponential gives
$$\displaystyle \begin{aligned} \left\langle \Delta T\right\rangle =\left(\frac{\tau}{q}\right)\varGamma\left[\frac{1}{q},\left(\frac{t_{1}}{\tau}\right)^{q}\right],{} \end{aligned} $$
(5.42)
where \(\varGamma \left [\left (1/q\right ),\left (t_{1}/\tau \right )^{q}\right ]\) is the upper incomplete gamma function.

5.6.8 Information Entropy and Stretched Exponential

The Shannon information entropy measures the amount of uncertainty in a given distribution, which is a measure of unpredictability. Similar to the power law distribution, the continuous information entropy of the stretched exponential distribution is defined as \(H\left [f\left (t\right )\right ] = \intop _{_{0}}^{T} f\left (t\right ) \log \left [1/f\left (t\right )\right ] dt\), where \(f\left (t\right )\) is the probability density function, \(f\left (t\right ) = \left (q/\tau \right ) \left (t/\tau \right )^{q-1} \exp \left [-\left (t/\tau \right )^{q}\right ]\), where \(\tau > 0\) is the relaxation time and \(0 < q < 1\) is the shape exponent and t is time. The \(\log \left [1/f\left (t\right )\right ]\) is the information content that an event will occur at time t with probability \(f\left (t\right )\), and, if \(f\left (t\right )\) is high, then knowledge that event occurred gives very little information, since it had a high probability of occurrence to start with. After integration, the information entropy for the stretched exponential distribution can be expressed as
$$\displaystyle \begin{aligned} H=\gamma\left(1-\frac{1}{q}\right)+\ln\left(\frac{\tau}{q}\right)+1,{} \end{aligned} $$
(5.43)
where \(\gamma =\) 0.57721 is the Euler-Mascheroni constant. For \(q = 1\), i.e. for the exponential distribution \(H = \ln \left (\tau \right ) + 1\), which is independent of q, and for \(\tau = 1\), \(H = 1.\)
The uncertainty is low for low q-values, and it reaches its maximum of \(\gamma +\ln \left (\tau /\gamma \right )\) at \(q = \gamma \), regardless of the relaxation time, and then slowly decays. For a given q, the uncertainty, or the unpredictability, increases with an increase in relaxation time, i.e. systems that are slow to relax are less predictable, see Fig. 5.26. Note that the uncertainty here refers to the forecasting ability of the \(f\left (t\right )\) and not to the uncertainty or the error in q.
Fig. 5.26
Information entropy vs. q for different \(\tau \)
Bild vergrößern

5.6.9 Non-stationary Poisson Process

In a non-stationary, also called a non-homogeneous, Poisson process, the underlying probability distribution and the intensity of the occurrence of events, \(I\left (t\right )\), vary with time. If \(\Lambda \left (t_{1},t_{1}+\Delta T\right ) = \intop _{t_{1}}^{t_{1}+\Delta T}I\left (t\right )dt\) is the expected number of events in the time interval \(\left (t_{1},t_{1}+\Delta T\right )\), then, according to the Poisson distribution, the probability of having exactly N events in that time interval is \(\Pr \left [N;\left (t_{1},t_{1}+\Delta T\right )\right ] = \left [\Lambda \left (t_{1},t_{1}+\Delta T\right )\right ]^{N} \exp \left [-\Lambda \left (t_{1},t_{1}+\Delta T\right )\right ] / N!\). The expected total number of events is \(\intop _{t_{0}}^{T}I\left (t\right )dt\), where \(t_{0}\) is the beginning and T is the end of the process. The probability of having zero events during that time is \(\Pr \left [N=0;\left (t_{1},t_{1}+\Delta T\right )\right ] = \exp \left [-\Lambda \left (t_{1},t_{1}+\Delta T\right )\right ]\), and the probability of having at least one event is
$$\displaystyle \begin{aligned} \Pr\left[N\geq1,\left(t_{1},t_{1}+\Delta T\right)\right]=1-\exp\left[-\Lambda\left(t_{1},t_{1}+\Delta T\right)\right].{} \end{aligned} $$
(5.44)
The non-stationary Poisson process can be applied to model the expected time and size distributions of seismic activity after step loading, e.g. after production blasts, pillar failures, or after larger events. If \(I\left (t,P\right )\) is the intensity function of the underlying process defined in such a way that \(\Lambda \left [\left (t_{1,}t_{1}+\Delta T\right );\left (P_{1},P_{2}\right )\right ] = \intop _{t_{1}}^{t_{1}+\Delta T}\intop _{P_{1}}^{P_{2}}I\left (t,P\right )dPdt\) is the expected number of events in the time interval \(\left (t_{1},t_{1}+\Delta T\right )\) and within the potency range \(\left (P_{1},P_{2}\right )\), then the probability of having exactly N events in this time interval and in this potency range is
$$\displaystyle \begin{aligned} \begin{array}{rcl} & &\displaystyle \Pr\left[N;\left(t_{1},t_{1}+\Delta T\right);\left(P_{1},P_{2}\right)\right]\\ & &\displaystyle =\frac{\Lambda\left[\left(t_{1},t_{1}+\Delta T\right);\left(P_{1},P_{2}\right)\right]^{N}}{N!}\exp\left\{ -\Lambda\left[\left(t_{1},t_{1}+\Delta T\right);\left(P_{1},P_{2}\right)\right]\right\} . \end{array} \end{aligned} $$
If within that time interval the time and size distributions are Independent, then \(I\left (t,P\right ) = I\left (t\right ) I\left (P\right )\) and \(\intop _{t_{1}}^{t_{1}+\Delta T} I\left (t\right ) dt = \Lambda \left (t_{1},t_{1}+\Delta T\right )\) is the expected number of events in this time interval that can be modelled by the Omori law or by the stretched exponential function. The integral \(\intop _{P_{1}}^{P_{2}} I\left (P\right ) dP = \Pr \left (P_{1},P_{2}\right )\) gives the probability of having an event in this potency range, and \(I\left (P\right )\) is the probability density function of the underlying size distribution, e.g. the upper truncated power law. Taking \(N = 0\) gives \(\Pr \left [N=0,\left (t_{1},t_{1}+\Delta T\right );\left (P_{1},P_{2}\right )\right ] = \exp \left [-\Lambda \left (t_{1},t_{1}+\Delta T\right )\Pr \left (P_{1},P_{2}\right )\right ]\), which is the probability that the first event in this potency range will occur after time \(t_{1}+\Delta T\). Therefore, the interval probability that at least one event will occur in potency range \(\left (P_{1},P_{2}\right )\) in this time interval is \(\Pr \left [N\geq 1;\left (t_{1},t_{1}+\Delta T\right );\left (P_{1},P_{2}\right )\right ] = 1 - \exp \left [-N\left (t_{1},t_{1}+\Delta T\right )\Pr \left (P_{1},P_{2}\right )\right ]\), which for \(P_{1} = P\) and \(P_{2} = P_{max}\) can be written as
$$\displaystyle \begin{aligned} \Pr\left[\geq P;\left(t_{1},t_{1}+\Delta T\right)\right]=1-\exp\left[-\Pr\left(\geq P\right)\cdot N\left(t_{1},t_{1}+\Delta T\right)\right].{} \end{aligned} $$
(5.45)
The probability \(\Pr \left (\geq P\right )\) can be given by the UT power law \(\Pr \left (\geq P\right ) = \left (P^{-\beta }-P_{max}^{-\beta }\right ) / \left (P_{min}^{-\beta }-P_{max}^{-\beta }\right )\), where \(\beta \) and \(P_{max}\) are parameters to be derived from data over the period \(\Delta t\) before the step loading.
Assuming the simple Omori relaxation function, i,e. \(c=0\), the number of events after the step loading is \(N\left (t_{1},t_{1}+\Delta T\right ) = k \left (t_{2}^{1-p}-t_{1}^{1-p}\right ) /\left (1-p\right )\) for \(p\neq 1\) and \(N\left (t_{1},t_{1}+\Delta T\right ) = \ln \left (t_{2}/t_{1}\right )\) for \(p=1\). For the stretched exponential relaxation function, the number of events after the step loading is given by Eq. (5.31). The \(N\left (t_{0},T\right )\) in Eq. (5.31) can be estimated in an on-line scheme as the aftershock data is being acquired, \(N\left (t_{0},T;t_{1}\right ) = N\left (t_{0},t_{1}\right ) / \left (1-\exp \left [-\left ((t_{1}-t_{0})/\tau \right )^{q}\right ]\right )\), where \(N\left (t_{0},t_{1}\right )\) is the observed number of events at time of forecast, \(t_{1} > t_{0}\), and \(\tau \) and q are estimated at the time \(t_{1}\) by the maximum likelihood method described above. This process is then repeated as the aftershock sequence progresses to firm up these estimates.
The total number of events to be produced by the sequence, \(N\left (t_{0},T\right )\), can also be estimated by the productivity law of aftershocks, \(N\left (t_{0},T\right ) = \alpha _{m} P_{m}^{\beta }\), where \(P_{m}\) is the known potency of the main shock, \(\alpha _{m}\) is the scaling factor to be calibrated, and \(\beta \) is the power law exponent.
Parameters k and p for the Omori relation and \(\tau \), q for the stretched exponential can be inverted on-line for the current ongoing sequence by the maximum likelihood method. They can also be taken a priori as average values from previous sequences in the area or, taking a Bayesian-like approach, a weighted average of the a priori and the on-line values (Reasenberg & Jones, 1989).
If after the step loading rock extraction is suspended, the relaxation process would end naturally at time \(t\left [N\left (t_{0},T\right )\right ]\). If, however, production continues or is restarted long before the relaxation process is completed, the stresses may accumulate and hazard may increase. In principle, mining could restart when seismic hazard dips below a predefined acceptable level and then stays there for a given period of time. This exclusion time would scale positively with seismic hazard before loading and the intensity of sudden loading which scales positively with the stress level in the area.

5.6.10 Aftershock Sequence After \(\log P=2.61\) Event

Here we will discuss the aftershock activity after the main shock (MS) with \(\log P=2.61\) that occurred on 07 July 2013, see description of the data up to the MS in Sect. 5.5. Figure 5.27 left shows the cumulative number of events with \(\log P\geq -2.0\) from 8 May to 23 July, where size of the event scales with the source volume and the colour indicates the distance of the event from the MS. The three vertical lines in red, blue, and green mark the times of the three largest events. The rate of events before the MS was steady with no increase in the rate or acceleration of activity before the MS. The first aftershock with \(\log P\geq -2.0\) was recorded 3.95, and the second 6.4 seconds after the MS. Within the first 16 hours after the main shock, there were 165 events that give the activity rate just over 10 events/hour, and within the first 171 hours, the rate of activity dropped to 1.4 events/hour.
Fig. 5.27
Cumulative number of events (left) and distances of events to the MS (right) vs. time
Bild vergrößern
Figure 5.27 right shows the distances from the MS during the same time with colour scaling with the logarithm of apparent stress, \(\log \sigma _{A} = \log \left (E/P\right )\). The bulk of the immediate aftershocks spread 670 metres from the MS. The largest of the immediate aftershock with \(\log P=0.15\) occurred 1.5 minutes after the MS and located 235 metres away. The largest event in the aftershock sequence with \(\log P=1.13\) occurred 162 hours after the MS and located 670 metres away.
Figure 5.28 left shows a 15 event moving window of the coefficient of variation of inter-event times, \(C_{v}\left (t\right )\), seismic diffusivity \(d_{s} = \left \langle d^{2}\right \rangle /\left \langle t\right \rangle \), and the seismicity rate change, where the colour scales with the mean distance of events in each moving window to the MS, that vary from 165 m in red to 391 m in blue. For details, see Sect. 5.6.2 in this chapter. The high \(C_{v}\left (t\right )\) indicates clustering of events in time. Seismic diffusivity, \(\left \langle d^{2}\right \rangle /\left \langle t\right \rangle \), increases due to an increase in the rate of seismic activity, i.e. decrease in the mean time between events, \(\left \langle t\right \rangle \), and/or increase in the spatial extent of events, i.e. increase in the mean distance from the MS or the mean inter-event distance \(\left \langle d^{2}\right \rangle \). Usually, \(d_{s}\) increases when a given system is overloaded and needs to relax. The seismicity rate change gives the probability that there is an increase or a decrease in the rate of events. Probability less than 0.5 means that there was a decrease in the event rate and above 0.5 that there was an increase. Parameter k specifies the magnitude of the change.
Fig. 5.28
Coefficient of variation of inter-event times (left), seismic diffusivity (centre), and seismicity rate change (right) before and after the MS
Bild vergrößern
The \(C_{v}\left (t\right )\) before the MS ranged between 0.79 and 1.4, and it spiked to 3.81 at the time of the MS indicating high clustering, but dropped to below 1.27 within 8 hours. The reference diffusivity started to increase 70 hours before MS and jumped to 3661.32 just after that and it took 63 hours to drop to pre-MS level.
Figure 5.28 right shows the probability that the rate of seismicity increased by at least 2 times with respect to the background level before the MS. The seismicity rate change jumped twice before the MS but reached a maximum 0.61 which is low. It spiked to 1.0 at the time of MS and stayed there for 17 hours dropping below the significant 0.75 level after 25 hours.
Figure 5.29 left shows the diffusivity vs. time during 147.7 hours before and 353.8 after an event with \(\log P=2.61\) shown here as a red circle. In this figure, colour scales with distance to the MS, and the size of the event here represents the radius of the source volume taken as a sphere, \(V=P/\Delta \epsilon \), where \(\Delta \epsilon \) is the assumed strain change at the source.
Fig. 5.29
Diffusivity vs. time of events with \( \log P \geq -2.0\) before and after the main shock (left). Scaling of \( \log \left \langle d \left (t \right )^{2} \right \rangle \) vs. \( \log t\) for 100 events before (in blue) and for 240 events after the main shock (in red) is shown on the right
Bild vergrößern
There was a steady rate of seismic activity of \(0.68\) events per hour before and a typical burst and a power law decay of aftershock activity after the main shock. Figure 5.29 right shows \(\log \left \langle d\left (t\right )^{2}\right \rangle \) vs. \(\log t\) scaling for 100 events before and for the first 240 events after the main shock. It shows a normal diffusion before with \(\gamma =1.04\) and a sub-diffusive process with \(\gamma =0.91\) after the MS.
Figure 5.30 first row shows the cumulative number of events 150 hours before and 200 after the MS with the stretched exponential (SE) fit to the first 16 and 176 hours of aftershock activity in red. Figure 5.30 second row shows the same but with the simple Omori fit (with \(c=0\)) in blue. In both cases, the forecast at the time \(t_{f}\) for the next 48 hours is shown in green. All statistics quoted on the right of the plots relate to the aftershocks at the time of forecast.
Fig. 5.30
Cumulative number of aftershocks and the SE fit in red (first row) and simple Omori in blue (second row) to the first 16 and 176 hours of aftershocks vs. relative time in hours. The forecasts for the next 48 hours are plotted in green. The probability of having a seismic event not smaller than \( \log P\) within 48 hours after the respective \(t_{f}\) by SE plotted in red and by simple Omori in blue, and the probabilities of having such an event during any 48 hours before the MS are plotted in black (bottom row)
Bild vergrößern
The SE fitted data well and provided a reasonable forecast at 16 and 171 hours after the MS. The SO fitted data at 16 hours well but failed in forecast and at 171 hours did not fit data well but provided a reasonable forecast. The coefficient of variation of time differences between events, \(C_{v}\left (t\right )\), increased from 1.75 in the first fit to 2.47 in the second fit due to the additional 76 intermittent events, but it shows that in both cases there is a temporal clustering and interdependence of events.
Figure 5.30 bottom row shows the interval probabilities, \(\Pr \!\left [\!\geq \!\log \! P;\!\!\left (t_{f},t_{f}{+}\Delta T\right )\!\right ]\), of having at least one event not smaller than a given \(\log P\) within \(\Delta T = 48\) hours after \(t_{f}\), calculated assuming SE, shown in red and SO in blue. The Poissonian probabilities during any 48 hours before the MS are shown in black. In general the SE provided better fit to data.
At 64 hours after the MS, the SE probabilities are considerably higher than before the MS, indicating that the relaxation process is still in progress and seismic hazard is high, and therefore, people should not enter the area affected by aftershock activity. For example \(\Pr \left [\log P\geq 0.0;64h\right ]\) is 0.4, while before the MS it was 0.265 and \(\Pr \left [\log P\geq 1.0;64h\right ]\) is 0.055, and before the MS it was 0.034. At 224 hours after the MS, the relaxation process is mainly completed and seismic hazard dropped. For example, \(\Pr \left [\log P\geq 0.0;240h\right ]\) is 0.079, while before the MS it was 0.265, and \(\Pr \left [\log P\geq 1.0;240h\right ]\) is 0.009, and before the MS it was 0.034.
Note that after the main shock mining was suspended and this is reflected in the interval probabilities, therefore here we compare seismic hazard before the MS driven by production with the interval probabilities after the MS with no production. Obviously, when production will resume, seismic hazard will increase, and therefore, the re-entry into the aftershock area needs to be done with caution.

5.6.11 Aftershock Sequence After \(\log P=5.47\) Event

This example describes aftershock sequence following a large \(\log P=5.47 \left (m_{HK}=4.6\right )\) and \(\log E=10.6\), mining triggered reverse slip event driven by tectonic sub-horizontal stress. The event was initiated off the bottom edge of a small mine and propagated over one kilometre away from the mine along sub-horizontal geological structure. Waveforms from the mine seismic system, from the close by mines, and regional stations were used to evaluate source parameters and mechanism of the main shock (MS) and aftershocks.
Figure 5.31 left shows three-component velocity waveforms recorded by 14 Hz sensors at distance 280 metres, Fig. 5.31 middle shows waveforms recorded by 4.5 Hz sensors at distance 4384 metres, and Fig. 5.31 right shows 200 seconds of waveforms recorded by broadband sensor at a distance of 92 km from the mine.
Fig. 5.31
Three-component velocity waveforms recorded by 14 Hz sensors at distance 280 metres (left) recorded by 4.5 Hz sensor at 4384 metres (middle) and by broadband sensor at distance 92 km from the mine (right)
Bild vergrößern
The moment tensor solution by the USGS, as well as location and mechanisms of aftershocks, indicates that the event represents a rupture along a plane with strike 240\(^{\circ }\) and dip 24\(^{\circ }\), in the mine’s coordinate system, which extends to the South-South-West of the mine. Analysis of waveforms indicates the two main episodes of rupture, the main shock and 0.7 seconds later the sub-event, that could well be treated as the first aftershock, see Fig. 5.31 left. This sub-event was most likely associated with the rupture of the fault segment about 950 m to the South-South-West from the primary main shock.
The spherical equivalent of the MS is shown in Fig. 5.32 by a large light grey circle that scales with the radius of the source volume with strain change \(\Delta \epsilon =\)10\(^{-4}\), \(r=\left [3P/\left (4\pi \Delta \epsilon \right )\right ]^{1/3}=856\) metres, where P is potency. However, the real shape of this event was rather closer to a flat ellipsoid span over the ellipse shown in Fig. 5.32 in blue. The event triggered 3670 aftershocks with \(\log P\geq -3.0\) within 156 days which are shown in Fig. 5.33. Most of the aftershocks located away from the mine are associated with the sub-horizontal planar structure approximated by ellipse. The largest aftershock with \(\log P=2.66\), shown in Fig. 5.32 in red, occurred within the mine 54 seconds after and 581 metres from the MS. The second largest aftershocks with \(\log P=2.54\) shown in Fig. 5.32 in orange occurred outside the mine 38 days after and 258 metres from the MS. The third largest aftershocks, with \(\log P=1.37\) occurred within the mine 15.6 hours after and 578 metres from the MS.
Fig. 5.32
Top (left) and section view (right) of the mine, including the nearest three mining operations. The main shock is shown in light grey and aftershocks coloured by the time of their occurrence
Bild vergrößern
Fig. 5.33
Cumulative number (left) and distances to the main shock (right) of seismic events with \( \log P \geq -3.0\) before and after the main shock
Bild vergrößern
Figure 5.33 left shows the cumulative number of events with \(\log P\geq -3.0\) during 1338.7 hours before and 3730.5 after the MS shown here as the large red circle. Colour here scales with distance to the MS, and the size of the event here represents the radius of the source volume taken as a sphere, \(V=P/\Delta \epsilon \), where \(\Delta \epsilon \) is the assumed strain change at the source, in this case \(10^{-2}\). There was a steady rate of seismic activity, \(\lambda _{B}=N\left (\log P\geq -3.0\right ) / \Delta t_{B} = 0.484\) per hour before the MS. Figure 5.33 right shows distances of seismic events to the MS over the same period of time where colour scales with the logarithm of apparent stress, \(\log \sigma _{A} = \log \left (E/P\right )\). The spatial distribution of events before the MS is mostly concentrated at the mine, at distances of 300 to 500 metres from the future MS, with only few events at distances over 500 metres. From the very beginning of the aftershock sequence, the spatial distribution of aftershocks extended to over 2000 metres from the MS. Approximately 2900 hours after the MS, one can see seismic activity at 200 metres from the MS associated with the rehab of mining operations.
Figure 5.34 top row shows the cumulative number of events with the stretched exponential (SE) fit at the time of forecast \(t_{f}=756\) and \(t_{f}=3057\) hours after the MS, shown in red, and its extrapolation into the next 48 hours in green. The simple Omori (SO), i.e. with \(c=0\), was also tested for these data sets and gave very similar results. All statistics quoted on the right of the plots relate to the aftershocks at the time of forecast. The first aftershock with \(\log P\geq -3.0\) was recorded 8 seconds after the MS.
Fig. 5.34
Cumulative number of aftershocks and the SE fit in red (first row) to the first 756 and 3057 hours of aftershocks vs. relative time in hours. The forecast for the next 48 hours is plotted in green. The probability of having a seismic event not smaller than \( \log P\) within 48 hours after the respective \(t_{f}\) by SE plotted in red and by simple Omori in blue, and the probabilities of having such an event during any 48 hours before the MS is plotted in black (second row)
Bild vergrößern
At \(t_{f}=756\) hours, the rate of aftershocks was \(\lambda _{A} = 2.84\), the background activity rate \(\lambda _{B}=0.48\). The coefficient of variation of time differences between events \(C_{v}\left (t\right ) = 1.6\), which indicates a degree of temporal clustering and interdependence of events. At that time, the mean distance of aftershocks to the MS was 857 metres, extending well beyond the boundary of the mine, which suggests the influence of local tectonics influenced by years of mining in the area.
Figure 5.34 bottom row shows the interval probabilities, \(\Pr \left [\geq \log P;\right .\)\(\left .\left (t_{f},t_{f}+\Delta T\right )\right ]\), of having at least one event not smaller than a given \(\log P\) within \(\Delta T = 48\) hours after \(t_{f}\), calculated assuming SE, shown in red, and SO in blue, which are practically the same.
The Poissonian probabilities during any 48 hours before the MS are shown in dark grey which, at \(t_{f}=756\), are considerably lower which implies that the relaxation process is still in progress and seismic hazard is high. For example, the SE probability, \(\Pr \left (\log P\geq 1.0;t_{f}+\Delta T=48\text{h}\right ) = 0.054\) as opposed to \(0.022\) for any 48 hours before the MS, which is 2.45 times higher. Note that the SE relaxation time, \(\tau \), at this stage is large, but it will get smaller as the relaxation process continues.
The forecast at \(t_{f}=3057\) hours gives a lower hourly rate of aftershocks of 1.13, and the activity rate over the last 1000 aftershocks is 0.51, which is very close to the premain shock background. The coefficient of variation of time differences between events increased to \(C_{v}\left (t\right ) = 2.29\) due to the intermittency of the recent aftershocks, which indicates a strong degree temporal clustering and interdependence of events. The spatial distribution of aftershocks extended even further beyond the boundary of the mine, which confirms the previous premise of tectonic or other mining influence. Figure 5.34 bottom right shows the interval probabilities, \(\Pr \left [\geq \log P;\left (t_{f},t_{f}+\Delta T\right )\right ]\), of having a seismic event not smaller than a given \(\log P\) within \(\Delta T = 48\) hours after \(t_{f}=3057\) hours, calculated assuming SE, shown in red, and SO in blue, which are practically the same and equal to the probabilities at any 48 hours before the MS.
At 756 hours, after the MS, the SE probabilities are considerably higher than before the MS, indicating that the relaxation process is still in progress and seismic hazard is high. For example, \(\Pr \left [\log P\geq 0.0;804h\right ]\) is 0.3, while before the MS it was 0.14 and \(\Pr \left [\log P\geq 1.0;64h\right ]\) is 0.053, and before the MS it was 0.022.
At 3105 hours after the MS, the relaxation process progressed to the stage that seismic hazard is at the same level it was before the MS.
Note that after the main shock mining was suspended, and this is reflected in the interval probabilities coming back to the pre-shock level rather quickly, and therefore, here we compare seismic hazard before the MS driven by external loading and production with the interval probabilities after the MS with no production. Obviously, when production will resume, seismic hazard will increase, and therefore the re-entry into the aftershock area needs to be done with caution.

5.7 Seismic Rock Mass Response to Blasting Sequence

5.7.1 Introduction

McGarr (1976b) showed that if after extracting a volume of rock \(V_{m}\), the altered stress and strain field can readjust to an equilibrium state through seismic movements only, the sum of seismic potency released within a given period of time would be proportional to the excavation closure, and in the long term \(\sum P = \sum V_{m}\). Then, assuming the seismic deformation will follow the open-ended power law with parameters \(\alpha \), \(\beta \), one can estimate the maximum possible event size \(P_{max}\) in the seismic sequence, \(P_{max} = \left [\left (1-\beta \right )V_{m}/\left (\alpha \beta \right )\right ]^{1/\left (1-\beta \right )}\). Similar equations were derived by Wyss (1973), Smith (1976), McGarr (1976a), and Molnar (1979) to estimate the expected maximum magnitude of earthquakes, and by McGarr (1984) for mines. See also section Balance of the Effective Volume Mined and \(P_{max}\)in Chap. 4.
Mendecki (2001) studied seismic rock mass response to extraction blasting in a long wall with four 30 m panels at a depth of 3000 m at Tau-Tona mine. The area was covered by a high-resolution microseismic network of five three-component accelerometers that provided enough data to gain insight into the excitation and relaxation processes associated with a sudden transient closure after production blasts. During five days of observations, 15 panels were blasted—minimum 3, maximum 6 panels per day, and no blasts on Sunday. In each stope, there were two rows of 0.9 m long blasting holes separated by 0.5 m vertically and horizontally. They were blasted in pairs top and bottom, with about 200 ms delay. Thus it took approximately 12 to 15 seconds to blast a panel, with face advance by about 0.8 m. A few thousand events were recorded over the five days of observations, but only 1086 with well-developed P- and S-wave signatures were processed and used in analysis. The larger seismic events occurred during Day1 and Day3 after the production blast of at least three contiguous panels. It was concluded that blasting a number of contiguous panels degrades stiffness of the surrounding rock more and for a longer period of time than if the same number of panels were blasted non-contiguously. The loss of stiffness makes the system more vulnerable to larger events.
In another study, Mendecki (2005) analysed the seismic rock mass response to production by blasting at Mponeng mine. The rate of mining \(V_{m}/\Delta t\) ranged from 3 to 33 m\(^{3}\)/day and did not correlate with seismic hazard. Although the highest rate of mining was where non-contiguous panels were blasted, the seismic hazard remained low and larger events occurred during the periods of contiguous blasting.
Martinsson and Torrman (2020) developed a Bayesian hierarchical model that captured the dynamic relationships between seismic activity and production rate, depth, and size of the orebodies. Testing on data from the LKAB Malmberget mine, they inferred that the seismic half-life, defined as the amount of time required for the activity to fall halfway to its steady state value, ranges from weeks up to 2 months for the cases considered. They also found that the effect of the weekly production on the induced seismicity depends exponentially on the average production size in the orebody. The seismic exposure term reveals a dependency on depth, and, for cases considered, an increase in depth by 100 m doubles the seismic activity.
de Beer (2022) tested correlations between production rates, taking into account ring and drawbell blasting and mucking, and seismic potency response in a block cave mine. He found that the rate of mucking is a long term driver of seismic potency release both above and below the undercut. Birch et al. (2024) conducted a similar analysis, fitting a simplified version of a generalised linear model to tons bogged and blasted in weekly intervals over two years at Oyu Tolgoi block cave mine. They found that during this period, which included the start of drawbell development and production bogging, weekly tonnes bogged had much stronger influence on seismicity than blasting.
Most of these papers focus on the seismic rock mass response to production and production rate. Here, we explore the influence of the proximity of group blasting on the intensity of seismic response as measured by the number of larger, potentially damaging seismic events during and after blasting.

5.7.2 Scaled Volume and Proximity Index of Blasts

The intensity of the seismic rock mass response to a single extraction, undercut, or preconditioning blast scales positively with the stress level at the blasting site and the surrounding area and with the volume of rock extracted. In the case of multiple blasts, it also depends on the number and the proximity of these blasts, i.e. the time differences and the distances between them. For a given stress condition and volumes of rock blasted, more clustered blasts result in a more intense seismic response.
Let us denote the volume of the first blast in a sequence of two blasts by \(V_{b1}\) and ask what should be the time delay, \(\Delta t_{21}\), and the space separation, \(\Delta d_{21}\), to the second blast to avoid superposition of seismic activity. First, let us define the scaled volume of the first blast as
$$\displaystyle \begin{aligned} V_{sb1}=\frac{V_{b1}}{\Delta t_{21}/t_{\alpha1}+\Delta d_{21}/d_{\alpha1}},{} \end{aligned} $$
(5.46)
where the numerator is the volume of rock blasted during the first blast, \(V_{b1}\), and the dimensionless denominator quantifies the proximity in time and in space between the first and the subsequent blast. The proximity in time is quantified by the ratio \(\Delta t_{21}/t_{\alpha 1}\), where \(t_{\alpha 1}\) is the characteristic seismic relaxation time, and the proximity in space is measured by the ratio \(\Delta d/d_{\alpha 1}\), where \(d_{\alpha 1}\) is the characteristic size, or the radius, of the seismic relaxation zone. Both \(t_{\alpha 1}\) and \(d_{\alpha 1}\) are functions of \(V_{b1}\). The best proxies for \(t_{\alpha 1}\) and \(d_{\alpha 1}\) are the calibrated exclusion or re-entry time, \(t_{r}\), and the characteristic size of the exclusion zone, \(d_{e}\). Alternatively, they can be scaled appropriately. For a given \(\Delta t_{21}\) and \(\Delta d_{21}\), the larger \(t_{\alpha 1}\) and/or \(d_{\alpha 1}\), the larger \(V_{sb1}\). Since the seismic rock mass response to blasting scales positively with the stress level, so would \(t_{\alpha 1}\) and \(d_{\alpha 1}\), and therefore, Eq. (5.46) indirectly caters for the stress level.
If the time difference between the first and the second blast \(\Delta t_{21}\) is equal to \(t_{\alpha 1}\) and the distance between them, \(\Delta d_{21}\), to \(d_{\alpha 1}\), then
$$\displaystyle \begin{aligned} V_{sb1}\left[\Delta t_{21}=t_{\alpha1},\Delta d_{21}=d_{\alpha1}\right]=V_{b}/2.{} \end{aligned} $$
(5.47)
Now, we can define the proximity index of the first blast to the second blast as the ratio,
$$\displaystyle \begin{aligned} I_{pb1}=\frac{V_{sb1}}{V_{sb1}\left[\Delta t_{21}=t_{\alpha1},\Delta d_{21}=d_{\alpha1}\right]}=\frac{2t_{\alpha1}d_{\alpha1}}{t_{\alpha1}\Delta d_{21}+d_{\alpha1}\Delta t_{21}}.{} \end{aligned} $$
(5.48)
If \(\Delta t_{21} = t_{\alpha 1}\) and \(\Delta d_{21} = d_{\alpha 1}\), then \(I_{pb1}=1\). For a given \(t_{\alpha 1}\) and \(d_{\alpha 1}\), the index \(I_{pb1}\) measures the deviation from the desired proximity of the second blast from the first one in time and/or in space. The case \(I_{pb1}\leq 1\) means the desirable outcome or better, and \(I_{pb1}>1\) means that, for a given \(V_{b1}\), the second blast may be too close in space and/or in time. Since \(t_{\alpha 1}\) and \(d_{\alpha 1}\) scale positively with \(V_{b1}\), the larger this volume the longer the relaxation time and the larger the relaxation zone. In principle both, \(t_{\alpha }\) and \(d_{\alpha }\) would scale with the cube root of the volume of rock blasted.
Equation (5.48) allows testing whether the preferred time difference and distance between two planned blasts are below the acceptable level of \(I_{pb}\) that produces a manageable seismic response.
In practice, mines frequently group more than two blasts to optimise the required exclusion time. The most practical way to test the proximity is to apply the above formulas to each two consecutive blasts. However, while all pairs of consecutive blasts may be reasonably separated in time and space, there may be some non-consecutive blasts missed by this scheme. For example, the case of three blasts carried out within a small \(\Delta t\) where the first and the second blasts are separated enough in space but the third one is very close in space to the first one. Here, if measured in relation to the second blast, the third one would give reasonable space separation resulting in low \(V_{sb}\) and \(I_{pb}\), while it would not be the case if the first blast would be measured in relation to the third one.
To detect such cases, one can consider all combinations of two blasts in a given blasting sequence. For \(n_{b}\) blasts, there are C\(_{2}^{n_{b}} = n\left (n-1\right )/2\) combinations of two blasts, and for each, we can calculate \(I_{pb}\). Then, we can sort \(I_{pb}\) from the largest to the smallest and compare with the sorted values of consecutive blasts to see if these two lists are the same.

5.7.3 Example: Proximity of Blasting and Seismic Response

The example below is based on the best blasting data available to the author for publication.
Proximity of Blasting
Figure 5.35 left shows the cumulative number of 3296 events with \(\log P_{min}\geq -3.0 \left (m\geq -1.1\right )\) over 91 days between 6 December and 7 March in an open stopping hard rock mine. In the figure, the size of the event scales with the source volume, and the colour indicates the distance of that event from the \(\log P=2.23 \left (m=2.4\right )\) main shock (MS) that occurred on 19 February at 13:09:48 at level of 1805. The second largest event with \(\log P=0.84 \left (m=1.5\right )\) occurred on 26 December at 14:16:12 at the 1706 level. The three vertical lines in red, blue, and green mark the times of the three largest events. There were 52 mid-size events with \(\log P\geq -1.0 \left (m\geq 0.25\right )\), which gives rate of 0.57/day. The mean recurrence interval, \(\bar {t}\left (\log P\geq -1.0\right ) = 34.82\) days with a standard deviation of \(35.86\) days, which gives the coefficient of variation \(C_{v} = 1.03\). There was no change of rate or acceleration of seismic activity before any of the three largest events. Figure 5.35 right shows the distances of events from the MS during the same time, with colour scaling with the logarithm of apparent stress, \(\log \sigma _{A} = \log \left (E/P\right )\). The bulk of these events locate between 200 and 400 metres from the MS, with the exception of the string of events 120 metres away that persisted for days after the MS.
Fig. 5.35
Cumulative number of events (left) and their distances to the MS (right) vs. time
Bild vergrößern
Figure 5.36 left shows the volume of rock blasted where the size of the rectangle scales with the volume of the blast with colour indicating the distance between consecutive blasts varying from 0 metres in red to 891 in blue. The left bottom of the rectangle marks the time of the blast. A timeline of seismic events and they sizes is shown at the bottom in grey. There were 82 blasts extracting 98240 m\(^{3}\) of rock, the smallest 48.5 m\(^{3}\) and the largest 5597.3 m\(^{3}\) at 1728 level, i.e. 77 metres below the MS. All 3321 combinations of two blasts were tested to confirm that they comply with the consecutive order.
Fig. 5.36
Cumulative number of events and volume of rock blasted (left). The cumulative volume of rock blasted with seismic events in light grey plotted at the bottom (right)
Bild vergrößern
Figure 5.36 right shows the cumulative plot of volume of rock blasted where colour scales with the distance between consecutive blasts. There was an increase in the rate of volume of rock blasted before the second largest event and even more so before the MS. There was also an increase in the number of closely spaced consecutive blasts before these two events.
Figure 5.37 shows the histogram of the volumes of rock blasted and the distances between consecutive blasts. There were 12 blasts with \(V_{b}>2000\) m\(^{3}\) and three blasts with \(V_{b}>3000\) m\(^{3}\), the maximum single blast was 5579 m\(^{3}\), and the minimum 49 m\(^{3}\). There were two pairs of blasts practically at the same spot but separated by 12 and 24 hours, respectively. There were three blasts at a distance of less than 10 m from their predecessor, 11\(<\)50 m, and 13\(<\)100 m, 15\(<\)150 m and 20 <200 m.
Fig. 5.37
Histograms of volumes of rock blasted (left) and distances between consecutive blasts (right)
Bild vergrößern
Figure 5.38 left shows the distances between consecutive blasts where colour scales with the distance to the MS. It also shows the cumulative volume of rock blasted in blue. Note the accelerating rock extraction before the MS.
Fig. 5.38
Distances between consecutive blasts with cumulative volume of rock blasted in blue (left) and the scaled volume of consecutive blasts (right). Seismic events in light grey vs. time are plotted at the bottom of both figures
Bild vergrößern
Since we do not have the calibrated values for the exclusion zone, we scaled it as, \(d_{ej} = 10\cdot S_{bj}\), where \(S_{bj} = V_{bj}^{1/3}\) is the characteristic size of rock blasted, which gave \(d_{emin} = 36\) metres for \(V_{bmin} = 48.5\) m\(^{3}\) and \(d_{emax} = 177\) metres for \(V_{bmin} = 5579\) m\(^{3}\). The two red horizontal lines in Fig. 5.38 left mark the minimum and the maximum radius of the scaled exclusion zone for all blasts.
We also do not have calibrated re-entry times, and therefore, we scaled them as follows:
$$\displaystyle \begin{aligned} t_{rj}=t_{rmin}+\left(t_{rmax}-t_{rmax}\exp\left[-\left(\left(S_{bj}-S_{bmin}\right)/S_{bmax}\right)^{q}\right]\right),{} \end{aligned} $$
(5.49)
where \(t_{rmin}\) and \(t_{rmax}\) are the assumed, for a given range of \(V_{b}\), minimum and maximum re-entry times, respectively, \(S_{bj}\) are the characteristic sizes of the rock blasted, and q is the exponent of the stretched exponential relaxation function. Assuming \(t_{rmin} = 4\) and \(t_{rmax} = 36\) hours, \(S_{bmin} =3.6\) m, \(S_{bmax} = 17.7\) m, and \(q=0.75\), Eq. (5.49) gives the minimum re-entry time of 4 hours for \(S_{bmin} =3.6\) m and 24.5 hours for \(S_{bmax} =17.7\) m.
Figure 5.38 right shows the scaled volume of consecutive blasts where the colour indicates distance between consecutive blasts.
Figure 5.39 left shows the proximity index of all blasts, \(I_{pb}\), where the colour indicates distance between consecutive blasts. The minimum \(I_{pb}\) is 0.087 and is associated with \(V_{b1} = 49\) m\(^{3}\), with time delay to the next blast \(\Delta t_{21} = 0\), but the distance to the next blast \(\Delta d_{21} = 836\) m. There are two such blasts in the catalogue, on 10 and 17 of February. The maximum \(I_{pb}\) is 43.1 and is associated with \(V_{b1} = 2773\) m\(^{3}\) with time delay to the next blast \(\Delta t_{21} = 0\) but with \(\Delta d_{21} = 7\) m. There were 15 blasts with \(I_{pb} \geq 1\), of which 5 were during 19 days before the second largest event with \(\log P=0.84\), and 5 during 9 days before the MS with \(\log P=2.23\). The blast with the maximum \(I_{pb}=43.1\) was carried out 6 minutes before the MS.
Fig. 5.39
Proximity indices, \(I_{pb}\), of blasts (left) and Cum\(I_{pb}\) vs. time (right). Seismic events in light grey vs. time are plotted at the bottom of both figures
Bild vergrößern
Figure 5.39 right shows the cumulative proximity index of blasts, Cum\(I_{pb}\), vs. time. There was an increase in the index before the MS that started with the blast on 10 February. Table 5.6 lists the proximity parameters of 16 blasts from 10 February before the MS. The data in columns in bold were used to calculate the proximity index of the given blast.
Table 5.6
Proximity parameters of 16 blasts before the MS
Month/Day/Time: \(t_{j}\) and \(t_{j+1}\)
\(V_{bj}\)
d\(_{MS}\)
\(V_{bj+1}\)
\(\Delta t_{j+1,j}\)
\(t_{rj}\)
\(\Delta d_{j+1,j}\)
\(d_{ej}\)
\(V_{sbj}\)
\(I_{pbj}\)
02/10/13:03 and 02/10/13:03
84
517
49
0
7.2
26
44
143.7
3.4
02/10/13:03 and 02/10/13:03
49
495
778
0
4.
836
37
2.1
0.09
02/10/13:03 and 02/10/13:03
778
458
471
0
16.3
84
92
850.5
2.19
02/10/13:03 and 02/11/13:03
471
409
1572
24
14.3
709
78
43.6
0.19
02/11/13:03 and 02/16/13:03
1572
467
1216
120
19.2
232
116
190.8
0.24
02/16/13:03 and 02/16/13:03
1216
313
452
0
18.1
346
107
375.2
0.62
02/16/13:03 and 02/17/13:03
452
86
84
24
14.1
588
77
48.2
0.21
02/17/13:03 and 02/17/13:03
84
517
49
0
7.2
26
44
142.8
3.38
02/17/13:03 and 02/17/13:03
49
495
778
0
4.0
836
37
2.1
0.09
02/17 13:03 and 02/18/13:03
778
458
1216
24
16.3
537
92
106.3
0.27
02/18/13:03 and 02/18/13:03
1216
313
656
0
18.1
303
107
428.6
0.71
02/18/13:03 and 02/18/13:03
656
345
1320
0
15.6
714
87
79.8
0.24
02/18/13:03 and 02/19/13:03
1320
409
226
24
18.5
47
110
765.9
1.16
02/19/13:03 and 02/19/13:03
226
440
2733
0
11.3
368
61
37.5
0.33
02/19/13:03 and 02/19/13:03
2733
86
2066
0
21.5
7
140
58893.5
43.1
02/19/13:03 and 02/20/01:03
2066
84
470
12
20.4
588
127
397
0.38
Seismic Response to Blasting Before and After the Main Shock
Figure 5.40 left shows the cumulative number of 100 events before and 250 after the MS as well as the time and volumes of rock blasted. Size of the event scales with the radius of source volume, and the colour indicates the distance of that event from the MS. In the figure, the size of the rectangle scales with the characteristic size of the blast, \(\left (V_{b}\right )^{1/3}\), and colour scales with the distance of that blast from MS.
Fig. 5.40
CumN and volumes of rock blasted (left) and distances (right) of 100 events before and 250 after the MS
Bild vergrößern
Note that not every blast or blast sequence triggered a seismic response. Let us look at the seven blast sequences from 72 hours before to 60 hours after the MS:
1.
Two blasts 72 hours before the MS. \(V_{b1} =\)1216 \(^{3}\) at distance 313 m away from the future MS and \(V_{b2} =\)452 m\(^{3}\) at distance of 86 m. Rock mass responded seismically to these two blasts.
 
2.
Three blasts 48 hours before the MS. \(V_{b1} =\) 84 m\(^{3}\) at distance 517 m, \(V_{b2} =\)49 m\(^{3}\) at 495 m, and \(V_{b3} =\)778 \(m^{3}\) at 458 m. No seismic response.
 
3.
Three blasts 24 hours before the MS. \(V_{b1} =\) 1216 m\(^{3}\) at distance 313 m, \(V_{b2} =\)656 m\(^{3}\) at 345 m, and \(V_{b3} =\)1320 \(m^{3}\) at 409 m. A very light seismic response.
 
4.
Three blasts 6 minutes before the MS. \(V_{b1} =\) 226 m\(^{3}\) at distance 440 m, \(V_{b2} =\)2733 m\(^{3}\) at 86 m, and \(V_{b3} =\)2066 \(m^{3}\) at 84 m. Strong seismic response that triggered \(\log P=2.23\) event and 95 aftershocks within the first 12 hours.
 
5.
Two blasts 12 hours after the MS. \(V_{b1} =\) 469 m\(^{3}\) at distance 517 m, \(V_{b2} =\)2778 m\(^{3}\) at 345 m. Relatively weak seismic response, considering the volume of the second blast and the buoyancy of the rock mass after the MS. The response was delayed by 1.5 hour and delivered 12 events during 1.2 hours. After these 12 events, the activity rate dropped.
 
6.
Two blasts 48 hours after the MS. \(V_{b1} =\) 606 m\(^{3}\) at distance 409 m, \(V_{b2} =\)2703 m\(^{3}\) at 440 m. Practically, no seismic response, just three very small events within 40 minutes.
 
7.
One blast 60 hours after the MS. \(V_{b1} =\) 867 m\(^{3}\) at distance 517 m. Practically, no seismic response, just a few very small events within 7 hours.
 
The fact that the rock mass responded seismically to blasts close to the future MS may suggest that this was an area of higher stress. Indeed, the apparent stress of the MS was 0.46 MPa, which is the highest in the data set. If correct, this would support the concept of tap-testing of the criticality of a system, i.e. how close is a given system to the critical point at which its behaviour would change qualitatively. Systems close to criticality are more excitable. The concept of tap-tests is that the immediate seismic response to blasting contains information about the general levels of stress.
There was a seismic flurry that started 35.5 hours after the MS and delivered 21 events within 1.5 hours. The bulk of these events located 120 metres from the MS, see Fig. 5.40 right that shows the distances of these events from the MS with colour scaling with the logarithm of apparent stress. It may be considered a spontaneous activity, unless it was the response to a blast that was not logged.
Figure 5.41 left shows a 15 event moving window of the probability that the rate of seismicity increased by at least 3 times with respect to the activity before the MS. The colour scales with the mean distance of events in each moving window to the MS that vary from 128 m in red to 612 m in blue. For details, see subsection 5.6.2 in this chapter. The probability jumped above the significant 0.75 level for 3 hours in reaction to seismic activity associated with two blasts 72 hours before the MS. These events were on average over 300 metres from the MS. It jumped to 1.0 at the time of the MS and stayed above 0.75 level for 10 hours. Then it dropped briefly and started to rise again 13.5 hours after the MS reacting to the 12 events after blasting. Practically, after the MS, the activity rate was 3 times higher than before the MS for 15 hours. The probability of activity rate change jumped above the 0.75 level again in response to what looks like a spontaneous flurry of small events that started 35.5 hours after the MS and stayed there for 2.8 hours.
Fig. 5.41
Probability of seismicity rate change (left) and reference seismic diffusivity (right) in a 15 event moving window
Bild vergrößern
Reference seismic diffusivity can be defined as the ratio of the mean distance squared between the reference location, e.g. the main shock or the blast or the point of injection in case of hydrofracturing, and seismic events to the mean time between these events, \(d_{sr}=\left \langle d^{2}\right \rangle /\left \langle t\right \rangle \). An increase in the seismic reference diffusivity signifies development of spatial correlation within the system. Figure 5.41 right shows a 15 event moving window of seismic diffusivity over the time of interest where colour scales with the mean distance of events in each moving window to the MS that vary from 128 m in red to 612 m in blue. The horizontal red line marks the mean reference diffusivity before the MS, \(d_{srmeanb}=0.87\). It ignored the light seismic response to the three blasts 24 hours after the MS mainly because they were relatively close to the MS. It started to increase 21.5 hours before the MS in reaction to both, a modest increase in the activity rate and larger distances of these events from the MS. It jumped at the time of the MS but dropped down to the pre-MS level 7 hours later. It increased modestly again for almost 10 hours 26 hours after the MS due to an increase in distances and in activity. The reference diffusivity dropped significantly during the flurry of activity 35 hours after the MS mainly due to the smaller distances from the MS and dropping in activity rate.
Final Comments
It is important to schedule blasting sequences to limit the overlap of seismic relaxation. If larger blasts are too frequent and/or too close to each other, i.e. if the next blast is still during the excitation phase of the previous one, it may push a larger and larger part of the system into the sub-critical stage where it is more sensitive to small stress perturbations. Such systems are characterised by periods of low seismic activity followed by small bursts of activity, and they are less time predictable. There is also the possibility that such blasting may induce a larger seismic event that would not have happened otherwise, as opposed to advancing the clock for events that are almost ready to be triggered.
It is recommended to use seismic tap-testing to guesstimate how close the system is to the critical point, at which its behaviour could change qualitatively. Systems close to criticality are more excitable. There are three ways to tap-test the system: (1) Self tap-testing, where one can analyse the immediate response to small event(s) within the system. (2) Self tap-testing, where one can analyse the immediate response to remote larger events. (3) Active tap-testing where one can analyse the immediate seismic response to small blasts carried out in selected areas.
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0 International License (http://​creativecommons.​org/​licenses/​by/​4.​0/​), which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter's Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the chapter's Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder.
download
DOWNLOAD
print
DRUCKEN
Titel
Time Distribution and Seismic Hazard
Verfasst von
Aleksander J. Mendecki
Copyright-Jahr
2025
DOI
https://doi.org/10.1007/978-3-031-93239-7_5
Zurück zum Zitat Aki, K. (1981). A probabilistic synthesis of precursory phenomena. In D. W. Simpson, & P. G. Richards (Eds.), Earthquake prediction: An international review, Maurice Eving (pp. 566–574). American Geophysical Union, Washington D.C.
Zurück zum Zitat Amitrano, D. (2003). Brittle-ductile transition and associated seismicity: Experimental and numerical studies and relationship with the b-value. Journal of Geophysical Research, 108 (B1)(2044), 1–15. http://​doi.​org/​10.​1029/​2001JB000680.
Zurück zum Zitat Barnsley, M. F. (1988). Fractals everywhere. Academic Press.
Zurück zum Zitat Birch, D., Rigby, A., Simanjuntak, K., Mardiansyah, F., & Bayar, E. (2024). Establishing relations between seismic activity and production. In 9th International Conference on Mass Mining, Kiruna Sweden (pp. 1634–1643).
Zurück zum Zitat Chen, C.-C., Wang, W.-C., Chang, Y.-F., Wu, Y.-M., & Lee, Y.-H. (2006). A correlation between the b-value and the fractal dimension from the aftershock sequence of the 1999 Chi-Chi, Taiwan, earthquake. Geophysical Journal International, 167(3), 1215–1219.CrossRef
Zurück zum Zitat Davis, P. M., Jackson, D. D., & Kagan, Y. Y. (1989). The longer it has been since the last earthquake, the longer the expected time till the next? Bulletin of the Seismological Society of America, 79(5), 1439–1456.CrossRef
Zurück zum Zitat de Beer, W. (2022). Quantitative correlation of production and seismic response in block caves. In Y. Potvin (Ed.), 5th International Conference on Block and Sublevel Caving, Perth (pp. 473–486).
Zurück zum Zitat Epstein, B., & Lomnitz, C. (1966). A model for the occurrence of large earthquakes. Nature, 211, 954–956. http://​doi.​org/​10.​1038/​211954b0.CrossRef
Zurück zum Zitat Haldane, J. B. S. (1955). The measurement of variation. Evolution, 9(4), 484. http://​doi.​org/​10.​2307/​2405484.CrossRef
Zurück zum Zitat Heath, J. P. (2006). Quantifying temporal variability in population abundances. Oikos, 115, 573–581. http://​doi.​org/​10.​1111/​j.​2006.​0030-1299.​15067.​x.CrossRef
Zurück zum Zitat Henderson, J. R., Barton, D. J., & Foulger, G. R. (1999). Fractal clustering of induced seismicity in The Geysers geothermal area, California. Geophysical Journal International, 139(2), 317–324. http://​doi.​org/​10.​1046/​j.​1365-246x.​1999.​00939.​x.CrossRef
Zurück zum Zitat Hirano, R. (1924). Study of the aftershocks of the Kwanto earthquake. Journal of the Meteorological Society of Japan, 77(2).
Zurück zum Zitat Hirata, T. (1989). A correlation between the b value and the fractal dimension of earthquakes. Journal of Geophysical Research, 94(B6), 7507–7514.CrossRef
Zurück zum Zitat Kijko, A. (1982). A modified form of the first Gumbel distribution: model for the occurrence of large earthquakes. Part 1 Derivation of distribution. Acta Geophysica Polonica, 30(4), 333–340.
Zurück zum Zitat King, G. (1983). The accommodation of large strains in the upper lithosphere of the earth and other solids by self-similar fault systems: The geometrical origin of b-value. Pure and Applied Geophysics, 121(5-6), 761–815. http://​doi.​org/​10.​1007/​BF02590182.CrossRef
Zurück zum Zitat Kisslinger, K. (1993). The stretched exponential function as an alternative model for aftershocks decay rate. Journal of Geophysical Research, 98(B2), 1913–1921.CrossRef
Zurück zum Zitat Kohlrausch, v. R. (1854). Theorie des elektrischen Ruckstandes in der Leidener Flasche. Annalem der Physik und Chemie (ed. J. C. Poggendorff), Leipzig, 91/1(56), 179–214, p. 198.
Zurück zum Zitat Kvalseth, T. O. (2016). Coefficient of variation: the second-order alternative. Journal of Applied Statistics. http://​doi.​org/​10.​1080/​02664763.​2016.​1174195.
Zurück zum Zitat Mandelbrot, B. B. (1967). How long is the coast of Britain? Statistical self-similarity and fractional dimension. Science, 156, 636–638.CrossRef
Zurück zum Zitat Marsan, D. (2003). Triggering of seismicity at short timescales following Californian earthquakes. Journal of Geophysical Research, 108(B5), 2266.CrossRef
Zurück zum Zitat Martinsson, J., & Torrman, W. (2020). Modelling the dynamic relationship between mining induced seismic activity and production rates, depth and size: A mine-wide hierarchical model. Pure and Applied Geophysics, 177, 2619–2639. http://​doi.​org/​10.​1007/​s00024-019-02378-y.CrossRef
Zurück zum Zitat McGarr, A. (1976a). Upper limit to earthquake size. Nature, 262(5567), 378–379.CrossRef
Zurück zum Zitat McGarr, A. (1976b). Seismic moments and volume changes. Journal of Geophysical Research, 81(8), 1487–1494.CrossRef
Zurück zum Zitat McGarr, A. (1984). Some applications of seismic source mechanism studies to assessing underground hazard. In N. C. Gay, & E. H. Wainwright (Eds.), Proceedings 1st International Symposium on Rockbursts and Seismicity in Mines, Johannesburg, South Africa (pp. 199–208). South African Institute of Mining and Metallurgy.
Zurück zum Zitat McGuire, R. K. (1977). Effects of uncertainty in seismicity on estimates of seismic hazard for the east coast of the United States. Bulletin of the Seismological Society of America, 67(3), 827–848.
Zurück zum Zitat Mendecki, A. J. (2001). Data-driven understanding of seismic rock mass response to mining: Keynote Address. In G. van Aswegen, R. J. Durrheim, & W. D. Ortlepp (Eds.), Proceedings 5th International Symposium on Rockbursts and Seismicity in Mines, Johannesburg, South Africa (pp. 1–9). South African Institute of Mining and Metallurgy.
Zurück zum Zitat Mendecki, A. J. (2005). Persistence of seismic rock mass response to mining. In Y. Potvin, & M. R. Hudyma (Eds.), Proceedings 6th International Symposium on Rockburst and Seismicity in Mines, Perth, Australia (pp. 97–105). Australian Centre for Geomechanics.
Zurück zum Zitat Mendecki, A. J. (2008). Forecasting seismic hazard in mines. In Y. Potvin, J. Carter, A. Diskin, & R. Jeffrey (Eds.), Proceedings 1st Southern Hemisphere International Rock Mechanics Symposium, Perth, Australia (pp. 55–69). Australian Centre for Geomechanics.
Zurück zum Zitat Molnar, P. (1979). Earthquake recurrence intervals and plate tectonics. Bulletin of the Seismological Society of America, 69(1), 115–133.CrossRef
Zurück zum Zitat Omori, F. (1893). On after-shocks, Tech. rep., College of Science Imperial University of Tokyo.
Zurück zum Zitat Omori, F. (1894). On the aftershocks of earthquakes. Journal of College Science, Imperial University of Tokyo, 7, 111–200.
Zurück zum Zitat Pearson, K. (1896). Mathematical contributions to the theory of evolution: Regression, heredity, and panmixia. The Philosophical Transactions of the Royal Society of London, A, 253–318.
Zurück zum Zitat Reasenberg, P. A., & Jones, L. M. (1989). Earthquake hazard after a mainshock in California. Science, 243, 1173–1176.CrossRef
Zurück zum Zitat Savage, J. C. (1994). Empirical earthquake probabilities from observed recurrence intervals. Bulletin of the Seismological Society of America, 84(1), 219–221.CrossRef
Zurück zum Zitat Shakal, A. F., & Willis, D. E. (1972). Estimated earthquake probabilities in the north Circum-Pacific area. Bulletin of the Seismological Society of America, 62(6), 1397–1410.CrossRef
Zurück zum Zitat Smith, S. W. (1976). Determination of maximum earthquake magnitude. Geophysical Research Letters, 3(6), 351–354.CrossRef
Zurück zum Zitat Sornette, D., & Knopoff, L. (1997). The paradox of the expected time until the next earthquake. Bulletin of the Seismological Society of America, 87(4), 789–798.CrossRef
Zurück zum Zitat Utsu, T. (1957). Magnitudes of earthquakes and occurrence of their aftershocks (in Japanese). Zisin, 2(10), 35–45.CrossRef
Zurück zum Zitat Vieira, F. M. C. C., Diering, D. H., & Durrheim, R. J. (2001). Methods to mine the ultra-deep tabular gold-bearing reefs of the Witwatersrand Basin, South Africa. In W. A. Hustrulid, & R. L. Bullock (Eds.), Underground Mining Methods: Engineering Fundamentals and International Case Studies (pp. 691–704). Society for Mining, Metallurgy and Exploration, Inc (SME).
Zurück zum Zitat Wang, J. (1994). On the correlation of observed Gutennberg-Richter’s b-value and Omori’s p-value for aftershocks. Bulletin of the Seismological Society of America, 84(6), 2008–2011.CrossRef
Zurück zum Zitat Williams, G., & Watts, D. C. (1970). Non-symmetrical dielectric relaxation behavior arising from a simple empirical decay function. Transactions of the Faraday Society, 66, 80–85. http://​doi.​org/​10.​1039/​TF9706600080.CrossRef
Zurück zum Zitat Wyss, M. (1973). Towards a physical understanding of the earthquake frequency distribution. Geophysical Journal of the Royal Astronomical Society, 31(4), 341–359.CrossRef
Zurück zum Zitat Wyss, M., Sammis, C. G., Nadeau, R. M., & Wiemer, S. (2004). Fractal dimension and b-value on creeping and locked patches of the San Andreas fault near Parkfield, California. Bulletin of the Seismological Society of America, 94(2), 410–421.CrossRef