Zum Inhalt

3. Monitoring Rock Mass Stability

  • 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 untersucht die grundlegenden Konzepte von Zeit, Ordnung, Unordnung und Stabilität im Zusammenhang mit der Überwachung der Gesteinsmassen. Es beginnt mit der Untersuchung der philosophischen und wissenschaftlichen Perspektiven auf Zeit und ihre Rolle in natürlichen Prozessen, wobei der Pfeil der Zeit und das Konzept der Entropie hervorgehoben werden. Der Text vertieft sich dann in die Prinzipien der Selbstorganisation und der Kritik und erklärt, wie diese Konzepte auf das Verhalten von Rockmassen angewendet werden. Darin wird die Rolle der seismischen Aktivität bei der Überwachung der Stabilität der Gesteinsmasse diskutiert, einschließlich der Verwendung statistischer Parameter wie seismische Potenz, Energie und Diffusivität. Das Kapitel präsentiert außerdem Fallstudien und praktische Beispiele, um die Anwendung dieser Konzepte in Bergbauumgebungen zu veranschaulichen. Abschließend werden die Implikationen dieser Erkenntnisse für das Management seismischer Gefahren und die Gewährleistung der Stabilität der Gesteinsmassen diskutiert. Die Leser werden ein tieferes Verständnis der komplexen Dynamik erlangen, die das Verhalten der Gesteinsmassen bestimmt, und der Werkzeuge, die zur Überwachung und Steuerung der Stabilität im Bergbau zur Verfügung stehen.

3.1 Note on Time, Order, Disorder, and Stability

In 1708, Gottfried Wilhelm von Leibniz said “Time and space are not things, but order of things”, (Wiener, 1951). Or, as Albert Einstein put it “Time and space are modes by which we think, and not condition in which we live” (Forsee, 1967). Therefore, time is nature’s way to keep everything from happening all at once, (Atiyah, 1988). And John Wheeler (1911–2008) quipped that “Space is what prevents everything happening to me”. In essence, anything is possible if it happens too fast to be detected.
Time, as incorporated in the basic laws of physics, e.g. by Newton, Maxwell, or Einstein, does not distinguish between past and future, all processes are time reversible, meaning that they can proceed backwards as well as forwards through time. In his excellent book, Eddington (1928) introduced the concept of the arrow of time. He writes that “cause and effect are closely bound up with arrow of time, i.e. the cause must precede the effect. Thus in primary physics, which knows nothing of time’s arrow, there is no discrimination of cause and effect, but events are connected by a symmetrical causal relation which is the same viewed from either end”. He then states that the future is associated with more randomness and the past with less randomness.1 The subject of arrow of time and irreversibility was thoroughly explored by the Nobel prize winner Ilya Prigogine in Prigogine (1980) and Prigogine (1997), where he states that irreversibility can no longer be identified with a mere appearance that would disappear if we had perfect knowledge. He used the concept of entropy to distinguish between reversible and irreversible processes. Only irreversible processes contribute to entropy production, e.g. chemical reactions, heat conduction, diffusion. The second law of thermodynamics then states that irreversible processes lead to a kind of “one-sidedness” of time. The positive time direction is associated with the increase of entropy. The law of entropy increase is simply a law of increasing disorganisation. In this process, the initial conditions are forgotten. As Prigogine (1997) speculates: “The big bang was an event associated with an instability within the medium that produced our universe. It marked the start of our universe but not the start of time. Although our universe has an age, the medium that produced our universe has none. Time has no beginning, and probably no end”.
On a macroscopic scale, all processes in nature are dissipative. Natural systems consist of a large number of elements which, at any given time, are not in the same state. Therefore, in order to accommodate the differences, a macroscopic system spontaneously generates local flow of energy and momentum in the form of localised small-scale events, in addition to those imposed by the external conditions. Close to equilibrium, the distribution of fluctuations is more or less random, the correlation time and the correlation length are short, and nonlinearities are mostly hidden. In other words, near equilibrium fluctuations are harmless. Away from equilibrium, the system is more susceptible to the action of intermittent intrinsic instabilities that, due to their nonlinear nature, are agents of spatial and temporal correlations. Here the distribution of fluctuations is broader than Gaussian, with slower, power law like, decays or with additional peaks, facilitating rare but larger events that dominate the behaviour of the system. The finite values of spatial and temporal correlations measure the distance from equilibrium, and as this distance grows, the influence of nonlinearities increases. When the spatial range of fluctuations increases, elements of the different parts of the system interact, and the system can generate and maintain a reproducible relationship among its distant parts. In this process of self-organisation, the system creates spatial and temporal patterns that are not directly imposed by external forces. When the range of correlations becomes comparable with the system size, the resulting coherence, or order, may influence its behaviour qualitatively, and the system may become critical and undergo bifurcation, i.e. transition to another state (Nicolis & Prigogine, 1977). The divergence of correlation length indicates that the details of the system are irrelevant to its critical behaviour.
In essence, the traditional classical science evolved around stability and certitudes, and probability was associated with ignorance, while in reality we experience fluctuations, instabilities, multiple choices, and limited predictability. Once instability is included, the meaning of “law of nature” changes, and they no longer express certitudes but rather possibilities and need to be formulated on the statistical level (Prigogine, 1996) or, as Wallerstein (1999) said, “Probability is the only scientific truth there is”.
Figure 3.1 left illustrates the stochastic analogue of bifurcation described by Nicolis and Prigogine (1989).
Fig. 3.1
Illustrations of stochastic analogue of bifurcation (left) and complexity vs. entropy (right)
Bild vergrößern
The system at time \(t_{1}\) is described by a sharp unimodal peak around its mean value \(\left \langle X\right \rangle \). In other words, the system is firmly in the current state, and the internal fluctuations and/or the external influences are weak. At time \(t_{2}\) due to increased strength of fluctuations and/or changes in boundary conditions, the range of spatial correlation increased, and the system has a flat distribution, exploring regions of phase space that are far from the most likely value, preparing for bifurcation, i.e. a qualitative change of its behaviour due to a small smooth change in the parameter values. At this stage, one would expect seismic activity to develop spatial correlations. For example, the seismic response to blasting would have far wider spatial range than at the time \(t_{1}\). This is a state of marginal stability. Such correlations can only arise in systems in which the regime of Poissonian fluctuation, or disorder, has been overcome, i.e. there is an increased coherence and associated lower dimensionality of the system. In a Poissonian regime, the inter-event times are distributed with the same probability density function, i.e. the process is memoryless, and the vanishing of the covariance in the joint probability distribution in two sub-volumes implies the complete absence of spatial correlations, i.e. no interactions between events. Such a state can therefore be considered as the prototype of disorder. At time \(t_{3}\), the system transitioned into a multi-hump regime beyond bifurcation and is more likely to be at \(\left \langle X_{2}\right \rangle \) than \(\left \langle X_{1}\right \rangle \).
In general, the approach to the critical point and the nature of the instability depend on the complexity of the system that depends, in turn, on the degree of order or disorder in the system. A highly entropic state is random and not so complex, and a low entropic state is not complex because of its regularities (Crutchfield & Young, 1989; Kaneko & Tsuda, 2001), see Fig. 3.1 right. Therefore one would expect the state of marginal stability, state at time \(t_{2}\), to represent high complexity where the system itself hesitates what to do next. For complexity to feature, the system needs to be open to interaction with the environment where there is a net flow of energy. Closed systems would proceed towards an ordered or disordered equilibrium state, which is not complex.
Mining is an open system where transitions between ordered and disordered states are driven by stress and strain gradients imposed by rock extraction and by resulting seismic and aseismic deformation. Moreover, mining is not a spontaneous process. Rock extraction takes place at a particular place, at a particular time, and at a particular rate which are all highly variable compared to the tectonic regime. This interferes with the process of self-organisation. The average rate of deformation induced by mining is at least two orders of magnitude greater than the average slip rate of tectonic plates. The bulk of seismic activity in mines starts with rock extraction, increases with the extraction ratio of the ore body, and stops rather quickly with cessation of mining. Larger events tend to occur after the extraction ratio, or the depth of mining, or both reach a certain level. The complexity of the rock mass response to mining is driven mainly by two competing spatio-temporal processes of excitation and relaxation. In terms of the stress-strain behaviour of a given volume of rock, it would be the mode switching and the balance between strain hardening and softening that drives complexity. Since softer, heterogeneous systems spend more time after the peak stress and dissipate more, they generate more entropy at the costs of complexity. Thus, it may be that the relative drop in complexity, due to increase in entropy, gives rise to lower seismic hazard and more visible alerts. Increase in disorder leads to slower transitions and to diffused instabilities, while highly homogeneous systems may crack in one go with little warning or precursory behaviour.
An important agent in the development of spatial and temporal correlations in highly stressed rock is seismic activity itself. By breaking numerous asperities, seismic events smooth the system, allowing transfer of stresses over larger distances and thus paving the way for even larger events. Disorder, on the other hand, plays a stabilising role and, to a degree, can be engineered into the system by scattered layout and/or by a scattered sequence of mining or blasting. Disordered directions of local stresses and a slower loading rate, i.e. slower rate of mining, may also play a role, since it promotes healing and thus stress roughening.
Assuming that large seismic events are those small ones that were not stopped soon enough, the one way to manage seismic hazard is to engineer a mine layout that, together with the natural structures, is “rough” enough to limit the extent of ruptures. It has been postulated that mining scenarios that introduce spatial heterogeneity, or roughness, may de-correlate the system and be less likely to develop larger dynamic instabilities (e.g. van Aswegen & Mendecki, 1999; Handley et al., 2000; Mendecki, 2001, Figure 8; Mendecki, 2005).
However, there is a degree of confusion regarding the role disorder plays in rock mass stability. Since order is prerequisite of human survival, the impulse to produce orderly arrangements is inbred by evolution. In common speech, order describes organisation, structural regularity, and is associated with stability. We can achieve more if we act in organised manner, and therefore, there is a tendency to design mining a layout with simple geometrical shapes with straight lines. When in the late 80s I argued for a disordered or scattered mining layout in South African gold mines, a senior mine executive replied “a long-wall is a long-wall is a long-wall”, which in an essence promoted straight lines. It took a few years and countless rockbursts to change that perception.
But rock subjected to excessive stress also ruptures along straight lines. Yes, in aseismic mines, an orderly design may facilitate better financial outcome, but in deep hard rock mines it is the order or the smoothness of the mine layout and geological structures that promote more frequent and more destructive rockbursts. Put simply, the entropy or disorder measures the degree to which energy is mixed up inside a system, and therefore, it scales positively with the quantity of energy no longer available to do physical work.
This introduction explains why this chapter is about the degree of rock mass stability rather than prediction or forecasting. Keilis-Borok (1994) described the earthquake generating part of the solid Earth as a hierarchical nonlinear dissipative system. The same can be said about a rock mass subjected to mining. It consists of a hierarchy of geological structures under load, random heterogeneities, patches of rock resisting deformation where stresses are increasing, patches of fractured rock where stresses are decreasing, and passive volumes of rock not influenced by loading. Such a system shows partial self-similarity, fractality, and self-organisation, and its parts may remain in a sub-critical state even after larger events.

3.2 Stability of Deformation and Stability of a System

The surface which bounds all stress states that correspond to elastic deformation in six-dimensional stress space is called the yield or damage surface. The evolution of the yield surface with continued deformation specifies the manner in which the material hardens or softens during the deformation. The direction of the inelastic strain increment beyond the yield surface is determined by the inelastic potential. The inelastic potential surface is always normal to the inelastic strain rate vector \(\dot {\epsilon }_{in}\). The associated flow or normality rule applies to the material where yield and potential surfaces coincide, that is when the inelastic strain rate vector is always normal to the yield surface. Under the associated flow with a smooth yield surface, one would expect a strain softening behaviour before the instability. This could be described by the classical instability criterion formulated by Hadamard in 1903 (Truesdell & Noll, 2004), and rediscovered by Pearson (1956) and Hill (1958),
$$\displaystyle \begin{aligned} \intop_{\delta V}\dot{\sigma}\cdot\dot{\epsilon}_{e}\cdot dV+\intop_{\delta V}\dot{\sigma}\cdot\dot{\epsilon}_{in}\cdot dV\begin{array}{ccc} > & 0 & \mbox{stable}\\ < & 0 & \mbox{unstable} \end{array},{} \end{aligned} $$
(3.1)
where \(\dot {\epsilon }_{e}\) and \(\dot {\epsilon }_{in}\) are elastic and inelastic strain rates, \(\dot {\sigma }\) is the rate of change of stress, and dV  the volume of interest. The second inequality in Eq. (3.1) states that for unstable deformation of dV  the inner product of the next increment of stress with the next increment of strain should be negative,
$$\displaystyle \begin{aligned} d\sigma d\epsilon=d\sigma d\epsilon_{e}+d\sigma d\epsilon_{in}<0. \end{aligned}$$
For the elastic strain increments \(d\sigma d\epsilon _{e}>0\), therefore the deformation will be unstable only when the inelastic term balances the elastic term, so that there is a net strain softening. Near the end of the hardening regime, however, almost all further strain increments are inelastic. As stress and strain are tensors, there are many different components that may be influential in causing the instability.
At some stage, the system may cease to deform in a homogeneous mode and undergoes bifurcation. In general, bifurcation refers to a qualitative change of the object under study due to a change of parameters on which the object depends. In this case, it is the non-uniqueness of the deformation path. A bifurcation mode may or may not be amplified by continued deformation past the bifurcation point. After the displacement corresponding to the bifurcation point, there may be more than one incremental displacement field that satisfies the equilibrium and boundary conditions. If \(\dot {\sigma }_{1}\) and \(\dot {\sigma }_{2}\) are the stress rate fields corresponding to two such solutions, with \(\dot {\epsilon }_{1}\) and \(\dot {\epsilon }_{2}\) the respective strain rates, then the necessary condition for bifurcation to occur is that there exists an incremental displacement field such that \(\intop _{V}\Delta \dot {\sigma }\Delta \dot {\epsilon } = 0\), where \(\Delta \dot {\sigma } = \dot {\sigma }_{1}\)- \(\dot {\sigma }_{2}\) and \(\Delta \dot {\epsilon } = \dot {\epsilon }_{1} - \dot {\epsilon }_{2}\).
If the bifurcation grows, it may be symmetrical or asymmetrical. A symmetrical bifurcation (e.g. barrelling of the specimen) is stable, so \(d\sigma d\epsilon >0\), i.e. the deformation is stable when it is easier to start deformation elsewhere rather than to pursue it where it has begun. Asymmetrical bifurcation, characteristic of materials that show frictional and dilatational responses to stress, results in strain localisation in single, conjugate, or quasi-periodic multiple shear bands (Rudnicki & Rice, 1975; Hobbs et al., 1990; Muhlhaus et al., 1992). The possibility of localisation arises when one or more stress components within a volume are able to decrease with increasing strain (Cundall, 1990). Here the deformation may be unstable. Alternatively, strain localisation can be caused by relative strain softening, where material within a localised strain zone hardens at a lower rate than the adjacent body of rock (Hobbs et al., 1990; Jessell & Lister, 1991). The localisation is driven by the release of strain energy from an unloading region outside the localisation zone. Strain localisation may occur under both strain softening and strain hardening material response and is promoted by non-associated flow and/or by the presence of vertices on the yield surface (Hobbs et al., 1990; Olson, 1992).
Fredrich et al. (1989) conducted the first laboratory study to quantify all the constitutive parameters for pressure-sensitive dilatant materials in both the brittle and semi-brittle regime. They measured an internal friction coefficient, a dilatancy factor, being the ratio between the increments of inelastic volume strain and the inelastic shear strain, and a hardening modulus for Carrera marble deformed at room temperature and at different confining pressures. At a confining pressure of 5 MPa, they showed a brittle failure mode, and at 40–190 MPa, semi-brittle failure was observed. In the brittle region, shear localisation was not evident until the sample was deformed well into the post-failure region. In the semi-brittle regime, with increasing confining pressure, the hardening modulus increased, whereas the friction coefficient, dilatancy factor, and Poisson ratio decreased. Thus, the bifurcation model predicts that shear localisation is inhibited in the semi-brittle regime as long as the rock continues to strain harden.
Wawersik et al. (1990) investigated localisation of deformation theoretically and in the laboratory. They found that deviation from normality, i.e. associated flow, promotes localisation before peak stress for plane-strain compression. Triaxial deformation was characterised by a broad peak followed by gradual strain softening, with breakdown instability occurring only well beyond the peak stress.
In the case where the nucleation zone is localised and is developing far from the free surface, the shear band(s) are constrained by both elastic and, to a limited extent, inelastic deformation in the less deformed rock surrounding the shear zone. In such a case, hardening of the system undergoing deformation is to be expected even if the material softens within the nucleation zone. The critical states of stability of a three-dimensional body can generally occur only if the compression stress is of the same order of magnitude as the shear modulus of the material (Bazant & Cedolin, 1991).
Whether or when the process of strain softening and/or strain localisation becomes unstable depends on the energy release and the dissipation inside the softening zone and on the exchange of mechanical energy between the softening zone and the ambient rock. Stress decrease associated with softening within the nucleation zone induces some elastic un-straining and redistribution of stresses in the ambient rock. Energy released by elastic rebound of the surrounding rock is fed into the softening region and accelerates its deformation. The amount of energy fed into the nucleation zone correlates positively with its size and with the strain rate or, rather, rate of un-straining in the ambient rock. As soon as an input of elastic energy exceeds the energy dissipation due to the inelastic processes in the growing nucleation zone, the system becomes unstable.
One should distinguish between the stability of deformation defined by the stress-strain response of the material and the stability of the system being deformed. For a given loading, the necessary condition for instability of a system, e.g. a tunnel, stope, mined out fault, or dyke, is that the volume of net softening needs to reach a critical size, which introduces a size scale into the instability criterion. The necessary condition for the dynamic instability is that this critical volume needs to be overloaded suddenly, which introduces a critical time scale.
There is no universal rule how to determine the critical length scale; however, one can try for a proxy to gauge the proximity of the system to the critical point. Some of the proxies are based on the assumption that the stability of the rock mass subjected to mining can be related to its stiffness, i.e. its ability to resist deformation with increasing stress. While the overall stiffness of the rock mass is being maintained, the seismic response to mining, measured by the cumulative inelastic co-seismic deformation, e.g., seismic potency P, is expected to be proportional to the effective volume mined, \(\sum P\sim V_{meff}\). As mining progresses, the overall stiffness of the rock mass is being degraded and the rate of potency release may increase. With further degradation in stiffness, the response may become nonlinear with accelerating potency release, frequently associated with an increase in activity rate, signifying potential for larger instability. The dynamics of such instability depends on the ratio of the stiffness of the potentially unstable volume of rock to the stiffness of the surrounding rock mass. The higher this ratio the more energy will be released per unit of inelastic deformation at the source.
Another guiding idea in the interpretation of seismic activity is the concept of self-organisation into critical state, i.e. a state at which the correlation length becomes comparable with the system size. Intermittently, the correlations length may reach or even exceed the system size creating conditions conducive for larger instabilities. Here one would expect an increase in the mean distance between consecutive events \(\left \langle d\right \rangle \). It is assumed that the growth of long-range correlations within the rock mass allows for progressively larger events to be generated. Zoller et al. (2001) tested the critical state concept for earthquakes in California in terms of the spatial correlation range and found a scaling relation \(\log R \sim 0.7~\mbox{m}\) between the main shock magnitude m and the critical region R. However, in mines, seismic activity follows rock extraction, and a scattered mining operation may obscure spatial correlation and influence the distribution of distances between events.

3.3 Seismic Softening and Accelerating Deformation

Seismic Softening
The growth of the deformation processes up to the point of instability is called nucleation. Breakdown instability will only take place once a quasi-static and/or quasi-dynamic inelastic deformation has occurred within the critical volume of rock. The entire nucleation process prior to overall instability includes aseismic creep (e.g. Dieterich, 1992; Ohnaka, 1992), sub-critical crack growth, and dynamic instabilities of local to small scale. Experimental data on sub-critical crack growth in synthetic quartz crystals indicates that this process starts only above certain stress levels, which could be of the order of 50% of the rupture stress (Darot & Gueguen, 1986). During sub-critical crack growth, crack advance occurs by discrete, individually dynamic events, but with slow average rupture velocity. Though individual microseismic events may be dynamic, their evolution can be modelled as a continuous, quasi-static process because they release only small amounts of stress or seismic potency compared to the breakdown instability. The larger the event the stronger its influence on the stress and strain environment in the area and the more likely it will affect both the time and the size of the next seismic event.
Since the development of the nucleation zone is associated with overall strain softening, sources of seismic events located within the nucleation zone would be, on average, of a slower or softer nature. This effect would be magnified in cases where the nucleation zone interfaces through the fractured zone with the opening. In the laboratory experiments, crack branching and distributed damage associated with the fracture are observed only if the strength variation in the source region is greater than the stress concentration (Labuz et al., 1985; Labuz et al., 1985; Cox & Paterson, 1990). Outside, or at the interface of the nucleation zone, one would expect seismic events of a harder or faster nature characterised by higher apparent stress. The overall size of the nucleation zone increases with the size of the potential instability (Scholz, 1990) and with the degree of inhomogeneity and decreases with the increase in the strain rate (Kato et al., 1992).
One can consider an interpretation of instability in terms of the relative stiffness of the components of the system—an instability will occur when the stiffness of the nucleation volume is equal to or exceeds the unloading stiffness of the surrounding rock, see Fig. 3.2 left. Stuart (1981) stated that, for earthquakes, the proximity to instability can be measured by the ratio of the stiffness of the fault zone to that of the elastic surroundings—stability decreases as this ratio approaches unity.
Fig. 3.2
Interaction of inelastic behaviour of the nucleation volume with the stiffness of the surrounding rock mass (left). Typical stress-strain diagram of a rock sample: stable sequence from \(t_{0}\) to \(t_{2}\) and an instability from \(t_{2}\) to \(t_{3}\)(right)
Bild vergrößern
The stress and the strain jumps at the point of instability, \(\Delta \sigma \) and \(\Delta \epsilon \), depend on the properties of the rock within and, to a certain extent, outside the nucleation volume. The stronger and more homogeneous the rock mass the higher the stress drop and the higher the ratio of the stress drop over strain drop associated with instability. In particular, instability is predicted much nearer the peak load for localised nucleation volumes and for states of deviatoric simple shear than for states of axisymmetric compression (Rudnicki, 1977).
Accelerating Deformation
It is the nature of rock fracture and friction that breakdown instability does not occur without some preceding phase of accelerating deformation (e.g. Hirata et al., 1987; (Rudnicki, 1988); Scholz, 1990 Mendecki, 1993; Dieterich, 1994). In some cases, this phase is detected by the seismic monitoring system and in other cases not.
An increase in the rate of co-seismic deformation before instability can be reflected by an increase in the rate of seismic events, an increase in the number of mid-size events or, for the same rate and sizes of events, by the softer nature of individual events, occurring within the nucleation volume. A small but statistically significant decrease in seismic activity (quiescence) has been observed at the beginning of the strain softening stage, followed by an increase just before the instability (Brady, 1977b; Brady, 1977a; Main et al., 1992).
Figure 3.2 right illustrates the acceleration of deformation preceding instability. It is assumed that the surrounding rock mass is stiffer than the nucleation zone and undergoes strain hardening at the time when the nucleation zone softens. As a result, for the same increments of loading stress, \(\sigma _{2}-\sigma _{1}\), and \(\sigma _{3}-\sigma _{2}\), the nucleation volume experiences considerably larger increments of displacement, \(u_{1}\), \(u_{2}\), \(u_{3}\), than the surrounding rock (see Rudnicki, 1988). The longer the surrounding rock maintains its stiffness the later and less dynamic is the instability.

3.4 Statistical Parameters of Co-seismic Deformation

Seismic events can routinely be quantified by the following four independent parameters derived from recorded waveforms: the time of the event, t, location, \(X=x,y,z\), seismic potency, P, and the radiated seismic energy, E. From seismic potency and energy, one can derive apparent stress, \(\sigma _{A} = E/P\), energy index, EI, i.e. the ratio of the observed radiated seismic energy of that event E, to the average energy \(\bar {E}(P) = 10^{d\log P+c}\) radiated by events of the observed potency P, for a given area of interest ((van Aswegen & Butler, 1993)), and apparent volume, \(V_{A} = \mu P^{2} / E\), (Mendecki, 1993).
From the four independent quantities derived from waveforms, we can derive a number of seismicity parameters related to co-seismic deformation and associated changes in the strain rate, stress, and rheology of the process.

3.4.1 Seismic, Strain, Strain Rate, Stress, and Stiffness

Brune (1968) calculated average earthquake slip rates for major fault zones by \(\left \langle u\right \rangle /\Delta t=\sum M/\left (\mu A_{T}\right )\) or \(\left \langle u\right \rangle /\Delta t=\sum P/\left (A_{T}\right )\), where M is seismic moment, P potency, \(A_{T}\) is the total area of the shear zone involved in seismic slip, and \(\mu \) is the assumed rigidity. Kostrov (1974) generalised Brune’s formula to get an average strain and strain rate produced by the n events randomly distributed within a volume \(\Delta V\) rather than along an individual fault, over time \(\Delta t\),
$$\displaystyle \begin{aligned} \epsilon_{s}(\Delta V,\Delta t)=\frac{\sum P}{2\Delta V}\quad \text{and}\quad \dot{\epsilon}_{s}(\Delta V,\Delta t)=\frac{\sum P}{2\Delta V\Delta t}.{} \end{aligned} $$
(3.2)
Then taking \(E = \sigma _{s}\dot {\epsilon }_{s}\Delta V\Delta t\), he defined the average seismic stress, \(\sigma _{s}\),
$$\displaystyle \begin{aligned} \sigma_{s}(\Delta V,\Delta t)=\frac{1}{\dot{\epsilon}_{s}\Delta V\Delta t}\sum E=\frac{2\sum E}{\sum P}.{} \end{aligned} $$
(3.3)
Seismic stiffness, \(K_{s}\), then can be defined as the ratio of seismic stress to seismic strain,
$$\displaystyle \begin{aligned} K_{s}(\Delta V,\Delta t)=\frac{\sigma_{s}}{\epsilon_{s}}=\frac{4\Delta V\sum E}{\left(\sum P\right)^{2}}.{} \end{aligned} $$
(3.4)
Seismic stiffness measures the rock mass ability to resist deformation with increasing stress.

3.4.2 Seismic Viscosity, Relaxation Time, and Deborah Number

Rock mass resistance to seismic deformation can also be measured by seismic viscosity, \(\eta _{s}\), in Pa\(\cdot \)s, defined as the ratio of seismic stress to seismic strain rate (Kostrov & Das, 1988; Mendecki, 1997),
$$\displaystyle \begin{aligned} \eta_{s}(\Delta V,\Delta t)=\frac{\sigma_{s}}{\dot{\epsilon}_{s}}=\frac{4\Delta V\Delta t\sum E}{\left(\sum P\right)^{2}}.{} \end{aligned} $$
(3.5)
The concept of seismic viscosity is similar to the fluid mechanics concept of turbulent or eddy viscosity. Unlike ordinary or molecular viscosity, eddy viscosity does not describe the physical properties of the medium but characterises the statistical properties of the flow. Therefore, it does not have to be constant but varies in time and space. For a fixed \(\Delta V\), seismic viscosity would increase during quiescence, due to the increase in \(\Delta t\) if all other components in Eq. (3.5) are constant, or during a sequence of higher than average apparent stress events. Consequently, for a fixed \(\Delta V\) and \(\Delta t\), viscosity would decrease during a sequence of low apparent stress, or softer, events. The inverse of viscosity is called fluidity. Lower seismic viscosity, or high fluidity, implies easier seismic inelastic deformation and greater stress transfer due to seismicity. The kinematic seismic viscosity is \(\nu _{s}\) = \(\eta _{s} / \rho \), in m\(^{2} /\)s, where \(\rho \) is rock density.
Seismic relaxation time quantifies the rate of change of stress during seismic deformation, \(\tau _{s} = \eta _{s}/\mu \), and it scales positively with the usefulness of past data in forecasting seismic deformation. The lower the relaxation time the shorter the time span of useful data.
Seismic Deborah number is defined as \(De_{s} = \tau _{s}/t_{f}\), where \(t_{f}\) is the time of observation, or the time scale of the process, which in practical applications can be replaced by the expected life span of a structure under consideration. Deborah number, like the relaxation time, may be interpreted as the ratio of elastic to viscous forces, with De going to infinity for a perfectly elastic medium.
The concept of Deborah number was presented by Marcus Reiner in his after dinner speech to the Fourth International Congress on Rheology at Brown University in August 1963, and then published in Reiner (1964). In the paper, Reiner quotes the Prophetess Deborah who in the Book of Judges proclaimed “The mountains flowed before the lord”. In his speech, Reiner said “Deborah knew two things. First, that the mountains flow, as everything flows. But, secondly, that they flowed before the Lord, and not before man, for the simple reason that man in his short lifetime cannot see them flowing, while the time of observation of God is infinite”. If the time of observation is long or the relaxation time of the material is short, then “fluid-like” behaviour is to be expected. Conversely if the relaxation time of the material is large, or the time of observation short, then the Deborah number is high and the material behaves, for all practical purposes, as a solid.
Knowing the relaxation time, one can use the Deborah number to evaluate the potential stability of pillars over time \(t_{f}\). The higher the seismic Deborah number the more stable it is. Crush pillars, on the other hand, should be designed for shorter relaxation time so that they would soften and drop off the load within the designed \(t_{f}\).

3.4.3 Seismic Diffusivity

A non-equilibrium process which is moving towards equilibrium at the rate governed by its distance from equilibrium is described by the diffusion equation
$$\displaystyle \begin{aligned} \frac{\partial\phi}{\partial t}=D\nabla^{2}\phi=D\left(\frac{\partial^{2}\phi}{\partial x^{2}}+\frac{\partial^{2}\phi}{\partial y^{2}}+\frac{\partial^{2}\phi}{\partial z^{2}}\right),{} \end{aligned} $$
(3.6)
where \(\partial \phi /\partial t\) is the change of \(\phi \) with time, D is diffusivity, and \(\nabla ^{2}\phi \) is called the divergence of gradient or Laplacian of \(\phi \). The solution in the 4D space of the initial value problem \(\phi (t=0,\boldsymbol {X})=\delta ^{3}(\boldsymbol {X})\), where \(\delta \) here is the Dirac delta function, is
$$\displaystyle \begin{aligned} \phi\left(t,\mathbf{X}\right)=\frac{1}{\left(4\pi Dt\right)^{3/2}}\exp\left(-\frac{{\mathbf{X}}^{2}}{4Dt}\right).{} \end{aligned} $$
(3.7)
If we compare with the standard Gaussian function, \(g\left (x\right ) = \exp \left (-x^{2}/\sigma ^{2}\right )\), then the standard deviation \(\sigma = \sqrt {4Dt}\), which points to the linear dependence of the variance with time, \(\sigma ^{2}\sim t\), or \(\left \langle d\left (t\right )^{2}\right \rangle \propto t\), where \(d\left (t\right )\) is the distance travelled over time t. This is the so-called normal diffusion. The diffusivity D has dimension \(m^{2}/s\) and is interpreted in terms of a characteristic distance of the process which varies only with the square root of time.
Many different experiments though reveal deviations from normal diffusion, in that diffusion is either faster or slower, which is termed an anomalous diffusion. A useful characterisation of the diffusion process is through the scaling of the mean square displacement with time, \(\left \langle d\left (t\right )^{2}\right \rangle \sim t^{\gamma }\), where \(\gamma \) is the scaling index. The case \(\gamma =1\) relates to the normal diffusion, and all other cases are termed anomalous. The cases \(\gamma >1\) form the family of super-diffusive processes, including the particular case \(\gamma =2\) which is called ballistic diffusion. The cases \(\gamma <1\) are the sub-diffusive processes. Plotting \(\log \left \langle d\left (t\right )^{2}\right \rangle \) vs. \(\log t\) is an experimental way to determine the type of diffusion. Figure 3.3 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. 3.3
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 3.3 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 \(\gamma =1.07\) and a sub-diffusive process with \(\gamma =0.91\) after the MS.
Mendecki (1997) defined seismic diffusivity as the ratio of the mean distance squared between the reference location, e.g. the main shock or the blast, the point of injection in case of hydrofracturing, and seismic events \(d_{sr}\), or between the consecutive sources of events, \(\left \langle d_{IE}^{2}\right \rangle \), to mean time between these events, \(\left \langle t\right \rangle \),
$$\displaystyle \begin{aligned} D_{s}\left(\Delta V,\Delta t\right)=\frac{\left\langle d^{2}\right\rangle }{\left\langle t\right\rangle }.{} \end{aligned} $$
(3.8)
When calculating distances, one can also include half of their source radii.

3.4.4 Seismic Schmidt Number

Similar to the turbulent Schmidt number, which is the ratio of eddy viscosity to eddy diffusivity, Mendecki (1997) defined the seismic Schmidt number as the ratio of kinematic seismic viscosity to seismic diffusivity,
$$\displaystyle \begin{aligned} Sc_{s}=\frac{\nu_{s}}{D_{s}}=\frac{4\Delta V\Delta t\left\langle t\right\rangle \sum E}{\rho\left\langle d^{2}\right\rangle \left(\sum P\right)^{2}}.{} \end{aligned} $$
(3.9)
Note that Eq. (3.9) encompasses directly all four independent parameters which describe seismicity, namely: \(\left \langle t\right \rangle \), \(\left \langle d\right \rangle \), \(\sum P\), and \(\sum E\). Looking at Eq. (3.9), the seismic Schmidt number would decrease with an increase in activity rate, with a drop in apparent stress of events and with an increase in the mean distance between the consecutive events that may signify an increase in the spatial correlation range. As mentioned before, the mean distance between the consecutive events in mines may not be the best proxy for growing correlation range. In some cases, we also observe a migration of seismic activity towards the future main shock which pushes the seismic Schmidt number up.

3.4.5 Shape Factor—Sphericity

To take into account the strain localisation that may promote localisation of seismic activity before instability, we can introduce a shape factor of seismicity. Shape factor is an index influenced by the shape of the object but independent of its size. Small or thin objects have a larger surface area compared to the volume, and this gives them a large ratio of surface to volume. As the size of an object increases, without changing shape, this ratio decreases. Wadell (1933) defined sphericity index as the surface area of a sphere of the same volume as the object, \(A_{s}\), divided by the actual surface area of the object, A, and therefore,
$$\displaystyle \begin{aligned} A_{s}^{3}=\left(4\pi r^{2}\right)^{3}=36\pi\left(\frac{4}{3}\pi r^{3}\right)^{2}=36\pi V^{2}\Rightarrow A_{s}=\left(36\pi V^{2}\right)^{1/3} \end{aligned}$$
$$\displaystyle \begin{aligned} =\pi^{1/3}\left(6V\right)^{2/3}, \Psi=A_{s}/A=\pi^{1/3}\left(6V\right)^{2/3}/A.{} \end{aligned} $$
(3.10)
Sphericity measures how closely the shape of an object resembles that of a perfect sphere. Sphericity of the sphere is 1 by definition, and any object which is not a sphere will have sphericity less than 1. The sphericity of a cube is 0.806, tetrahedron 0.671, the common salt is 0.84, crushed coal 0.75, and crushed glass 0.65. To estimate the sphericity, \(\Psi \), of a set of seismic events given by their locations, we calculate the volume and the area of the smallest convex polygon that encloses these events, which is called a convex hull.

3.5 Seismic Stability Analysis

3.5.1 Assumptions and Data Selection

The seismic stability analysis has been developed over the last 30 years. The first formulation was in Mendecki (1993), where Eq. (3.1) was quoted, followed by Mendecki (1997) and numerous applications and modifications, e.g. van Aswegen et al. (1997), van Aswegen (2005), Rebuli and van Aswegen (2013).
The objective of seismic stability analysis is not to predict instability or to manage seismic exposure in the short term, but to guide control measures to mitigate seismic hazard. There is a view that these control measures just delay the inevitable. However, experience shows that scattered and slower rock extraction changes the nature of seismic release by producing more smaller or mid-size events and fewer large ones.
Theoretical considerations, laboratory studies, and seismic observations suggest that an inhomogeneous rock mass subjected to loading shares certain seismic symptoms when approaching instability:
  • An overall softening, measured by a decrease in the average value of the apparent stress, \(\sigma _{A}\), or energy index, EI.
  • Increased rate or accelerating deformation, measured by an increase in the activity rate, \(\lambda =1/\left \langle t\right \rangle \). Or, for the same activity rate, by lower apparent stress or larger apparent volume events that would result in increased rates of cumulative potency, CumP, or cumulative apparent volume, Cum\(V_{A}\).
  • Decrease in the dimensionality, or localisation, of seismic activity, measured by the fractal dimension, correlation dimension, or by the sphericity index of seismic sequences, \(\Psi \).
  • Increase in correlation length. In this case, one would expect to observe an increase in the spatial distribution of seismic activity and, in some cases, an increase in the number of mid-size events in different parts of the system. One could also observe an extended spatial distribution of the immediate seismic response to blasting or to mid-size events.
The following ratios depict the expected qualitative changes preceding larger instabilities:
$$\displaystyle \begin{aligned} \begin{array}{rcl} \Psi\mbox{\mbox{{$\frac{\mbox{apparent stress}}{\mbox{activity rate}}$}} \text{or} {$\Psi\frac{\mbox{apparent stress}}{\mbox{apparent volume}}$}} & = &\displaystyle \frac{\searrow}{\nearrow}=\,\Downarrow \\ \text{{$\Psi\frac{\text{seismic stress}}{\text{seismic strain rate}}$}}\;\text{or}\;\Psi\frac{\text{seismic viscosity}}{\text{seismic diffusivity}} & = &\displaystyle \frac{\searrow}{\nearrow}=\,\Downarrow.{} \end{array} \end{aligned} $$
(3.11)
All statistical parameters of co-seismic deformation described above directly or indirectly include \(\Delta V\) that, in the case of stability analysis, should be the seismogenic volume of rock that generates the next larger event. This is the volume we suppose to select seismic events from to plot a given stability function. However, a seismogenic volume is an expression which is not well defined. If we define it as a volume of rock that includes all interdependent events surrounding the future main shock, then it is a function of space and will be different at different locations. It will also scale positively with the size of the future main shock.
A stability analysis yields plots of a given stability function vs. time for a given site, and having done it for a number of sites, one can contour it on a plane or in space. The values of the stability function are calculated in a moving window. The window can be fixed in time or with a fixed number of events, in which case a variable in time. One way to select events into the moving window is to take what is available in the database. In this case, there may be events far away from the site that may have no influence on stability, or that may obscure the analysis. Alternatively, we may select a polygon based on past experience. Another option is to select events that exceed a given threshold of peak ground velocity, PGV , at the selected site. This influence based selection is more general, because if the threshold is set low it gives the first option and it facilitates automatic plotting.

3.5.2 Stability Example

Data
We analysed the last 60 days of seismic history before a \(\log P=2.61\) event at a mine here referred to as MineD, see the seismic hazard case study described in Chapter 5. Figure 3.4 left shows the convex hull span over all 6073 events \(\log P\geq -4.5\) available for analysis.
Fig. 3.4
Convex hull span over all available 6073 events with sites S1, S2, and S3 (left) and over 3578 events that generated \(PGV \geq 10^{-7}\) m/s at site S3 (right)
Bild vergrößern
The size of the event scales with the radius of source volume, and the colour indicates the time of the event, from the earliest in blue to the latest in red. Stability was assessed at three sites: at the source of the largest event S1, at the source of the second largest event S2, and at the centre of all 6073 events available for analysis S3. Distances between these sites are: S1 and S2 = 273 m, S1 and S3 = 185 m, and S2 and S3 = 118 m. Figure 3.4 right shows the convex hull span over 3578 events that generated \(PGV \geq 10^{-7}\) m/s at the centre of all 6073 events, shown as S3 in blue.
Distances of these events to site S3 range between 12 and 542 m. Tables in these figures give the number of events, \(N_{E}\), activity rate/day, the \(\log P\) range, the volume, \(V_{CH}\), and the sphericity index, \(\Psi _{CH}\), of the convex hull, the total volume of seismic sources with strain change = 0.0001, \(V\left (\Delta \epsilon \geq 10^{-4}\right )\), and the ratio of \(V_{CH}/V\left (\Delta \epsilon \geq 10^{-4}\right )\). After the selection of the 3578 events \(V_{CH}/V\left (\Delta \epsilon \geq 10^{-4}\right )\), ratio dropped from 177.2 to 8.85, which indicates that the volume selected for stability is reasonably saturated with co-seismic inelastic deformation. The sphericity index, \(\Psi _{CH}\), here dropped slightly from 0.86 to 0.83, however, while running moving windows through these 3578 events it varied between 0.61 and 0.86.
Figure 3.5 shows the cumulative number of events, CumN, and the cumulative apparent volume, Cum\(V_{A}\) in km\(^{3}\), for all 6073 events. The size of the event represents the radius of the source volume and colour scales with distance to the site S3. There were 17 mid-size events with \(\log P\geq 0.0\), which gives an activity rate 0.28/day, the coefficient of variation of these 17 events \(C_{v}\) is 0.87, but of all 6073 events \(C_{v}=1.32\), which indicates a degree of time clustering for the latter.
Fig. 3.5
CumN(left) and Cum\(V_{A}\)(right) vs. time plotted for all 6073 events
Bild vergrößern
The CumN vs. time plot is quite steady with almost constant activity rate, while Cum\(V_{A}\) shows a bit more structure and a significant increase in rate before the main shock. The largest event with \(\log P=2.61\) occurred on 07 July and located 185 m from site S3, the second largest with \(\log P=1.24\) on 14 June located 118 m away, and the third largest with \(\log P=1.06\) on 08 June 153 m away. The distance between the largest event and the second largest is 273 m, the largest and the third largest 85 m, and the second and the third largest 211 m.
Results
We tested a few different seismic stability functions defined by the combination of statistical parameters of co-seismic deformation described above, and all of them dropped to a low level before the \(\log P=2.61\) main shock. Some of them also dropped before a few mid-size events preceding the main shock. All plots below show the moving window-based qualitative history of a given stability function in black and, as a reference, the Cum\(V_{A}\) of the selected 3578 events marked by their sizes and colour indicating distances to site S3. The vertical axes are for Cum\(V_{A}\) in km\(^{3}\). We use a moving window of 119 events, which is 2 days of seismic activity of the selected events. All elements of the stability functions were normalised between 1 and 10.
Note that the Cum\(V_{A}\) vs. time plot of the 3578 events does not show the same change in rate before the main shock as that shown by the 6073 events. This increase is caused by three little clusters to the right of the main shock (see Fig. 3.4 left), and these events were too small and/or too far to generate \(PGV \geq 10^{-7}\) at Site 3 and were automatically excluded.
Figure 3.6 shows two seismic stability functions derived for site S3. The vertical red, blue, and green lines indicate times of the three largest events respectively. (1) \(\Psi _{CH}\cdot \text{Median}\left [\log \sigma _{A}\right ]/\lambda \), where \(\sigma _{A}\) apparent stress, \(\lambda \) the activity rate, and \(\Psi _{CH}\) sphericity of the convex hull span over 119 events in each moving window. This function dropped to a low level before all three largest events. (2) \(\Psi _{CH} \cdot \text{Median}\left [\log \sigma _{A}\right ]/\left (\lambda \cdot \sum u_{CAD}\right )\), where \(\sum u_{CAD}\) is the sum of the cumulative absolute displacements, CAD, in a given time window. CAD is derived from the GMPE for a given mine or area. This function also dropped to a low level before all three largest events.
Fig. 3.6
Time history of the first two seismic stability functions, in black, with Cum\(V_{A}\) of the selected data set
Bild vergrößern
Figure below shows the main components of these two stability functions.
Figure 3.7 left shows the time history of the \(\text{Median}\left [\log \sigma _{A}\right ]\) that drops convincingly, as desired in terms of the stability criteria described above, before the main shock, less so before the second and third largest events. Figure 3.7 right shows the time history of the activity rate, \(\lambda \), that is at high level before all three largest events and corrects the undesired increases in the \(\text{Median}\left [\log \sigma _{A}\right ]\).
Fig. 3.7
Time history of \(\text{Median}[ \log \sigma _{A}]\)(left), and the activity rate, \(\lambda \), (right)
Bild vergrößern
Figure 3.8 left shows the time history of the sphericity of the convex hull, \(\Psi _{CH}\), that is low before the second and third largest events, not so before the largest one. Figure 3.8 right shows the time history of the \(\sum u_{CAD}\) that is undesirably low before the largest event but complies with the assumptions before the second and third ones.
Fig. 3.8
Time history of sphericity, \(\Psi \)(left), and the sum of the cumulative absolute displacements, \( \sum u_{CAD}\)(right)
Bild vergrößern
Figure 3.9 left shows a Schmidt number based stability function for site S3, \(\Psi _{CH}\cdot \log Sc_{s}\left (V_{CH}\right )\), defined by Eq. (3.9) with \(\Delta V\) taken as the volume of convex hull, \(V_{CH}\). This function indicates low stability before all three largest events. Figure 3.9 right shows seismic diffusivity that is increasing, as assumed in Eq. (3.8), before the largest and the third largest events, no so before the second largest though.
Fig. 3.9
Time history of the Schmidt number (left) and diffusivity (right)
Bild vergrößern
Comparison Between Sites
The stability analysis described above was applied to site S3. Figure 3.10 below shows the same stability functions applied to sites S1, S2 and again for S3, for completeness.
Fig. 3.10
Time histories of the same three stability functions for site S1 (top row), S2 (middle), and S3 (bottom)
Bild vergrößern
Each site has a different set of events that generated \(PGV \geq 10^{-7}\), with different distances to a given site. They also have different activity rates: 31.9 events/day at sites S1, 43.72 at site S2, and 59.73 at site S3, and therefore, they have a different length of the moving window.
In this case study, there are no significant differences in qualitative behaviour of stability functions at these three sites, and they are all at a reasonably low level before the three largest events. The main reason is that all three sites are located relatively close to each other and also close to the main volume of seismic activity. In most cases studied to date stability function are reasonably similar for sites at distances less than 200. The notable exception would be if there is a flurry of small events close to one site that does not exceed the assumed PGV  threshold at the other site.
Somewhat surprisingly, in this case, sites S2 and S3 behave reasonably similar and gave better indication of instability before the largest event than site S1 that was deliberately located at site S1. One possible explanation is that site S1 attracted the lowest number of events. In general, the described stability analysis, or any other seismic alerts, performs better with higher activity rate.

3.5.3 Stability—General Comments

When analysing seismic stability, one should take into account the following limitations:
1.
Seismic stability analysis is based on theoretical considerations, laboratory experiments, and case studies. Therefore, it will always be influenced by local factors associated with a particular way of mining and by geological conditions.
 
2.
The utility of seismic stability analysis is in guiding control measures to mitigate seismic hazard. If the value of a given stability function is systematically dropping or stays low in a given area, then the mine may change the spatial and temporal manner of rock extraction, e.g. scatter the production blasts and/or slow down the rate of mining. Alternatively, they may try to trigger the potential seismic event, while people are not in the area.
 
3.
Events induced by production blasts may not be preceded by instability indicators. On the upside, people are not in the area during blasting. Production blasting with associated increase in seismic activity and strain softening will push the stability function lower. In a reasonably stable system, it should, however, recover relatively quickly. A simple stability function based on median apparent stress and seismic activity could be useful to monitor for safe re-entry.
 
4.
At any particular mine, tests should be carried out to select the most suitable stability function or functions, and the reference strain needs to be calibrated for a given GMPE at different sites.
 
5.
In some mining scenarios, we observe the migration of seismic activity towards the source of the future main shock. In these cases, the distances between consecutive events, \(\left \langle d^{2}\right \rangle \), may decrease before the main shock. If this is the expected outcome, one constructs the appropriate stability function.
 
6.
The seismic Schmidt number has four independent parameters which describe seismicity and, in theory, should be the most suited for stability analysis. In simpler mining scenarios, it is the case, however, in more complex environment, one or two parameters fail to comply with our assumptions and it may give inferior results.
 
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
Monitoring Rock Mass Stability
Verfasst von
Aleksander J. Mendecki
Copyright-Jahr
2025
DOI
https://doi.org/10.1007/978-3-031-93239-7_3
1
Without any mystic appeal to consciousness, it is possible to find a direction of time on the four-dimensional map by a study of organisation. Let us draw an arrow arbitrarily. If as we follow the arrow, we find more and more of the random element in the state of the world, then the arrow is pointed towards the future; if the random element decreases, the arrow points towards the past. That is the only distinction known to physics. This follows at once if our fundamental contention is admitted that the introduction of randomness is the only thing which cannot be undone.
 
Zurück zum Zitat Atiyah, M. (1988). Collected papers, Oxford.
Zurück zum Zitat Bazant, Z. P., & Cedolin, L. (1991). Stability of structures: Elastic, inelastic, fracture, and damage theories (1st edn.). Oxford University Press.
Zurück zum Zitat Brady, B. (1977a). An investigation of the scale invariant properties of failure. International Journal of Rock Mechanics & Mining Sciences, 14, 121–126.CrossRef
Zurück zum Zitat Brady, B. T. (1977b). Anomalous seismicity prior to rock bursts: implications for earthquake prediction. In M. Wyss (Ed.), Stress in the Earth (vol. 115, pp. 357–374). Pure Applied Geophysics.
Zurück zum Zitat Brune, J. N. (1968). Seismic moment, seismicity and rate of slip along major fault zones. Journal of Geophysical Research, 73(2), 777–784.CrossRef
Zurück zum Zitat Cox, S. J. D., & Paterson, L. (1990). Damage development during rupture of heterogeneous brittle materials: a numerical study. In R. J. Knipe, & E. H. Rutter (Eds.), Deformation mechanisms, rheology and tectonics (vol. 54, pp. 57–62). Geological Society, Special Publication.
Zurück zum Zitat Crutchfield, J. P., & Young, K. (1989). Inferring statistical complexity. Physical Review Letters, 63, 105–108.CrossRef
Zurück zum Zitat Cundall, P. A. (1990). Numerical modelling of jointed and faulted rock. In H. P. Rossmanith (Ed.), Mechanics of jointed and faulted rock (pp. 11–18). Balkema, Rotterdam.
Zurück zum Zitat Darot, M., & Gueguen, Y. (1986). Slow crack growth in minerals and rocks: Theory and experiments. Pure and Applied Geophysics, 124, 677–692.CrossRef
Zurück zum Zitat Dieterich, J. (1994). A constitutive law for rate of earthquake production and its application to earthquake clustering. Journal of Geophysical Research, 99(B2), 2601–2618.CrossRef
Zurück zum Zitat Dieterich, J. H. (1992). Earthquake nucleation on faults with rate- and state-dependent strength. Tectonophysics, 211(1-4), 115–134.CrossRef
Zurück zum Zitat Eddington, A. S. (1928). The nature of the physical world. University Press, Cambridge.
Zurück zum Zitat Forsee, A. (1967). Albert Einstein: Theoretical physicist. The Macmillan Company.
Zurück zum Zitat Fredrich, J., Evans, B., & Wong, T. F. (1989). Micromechanics of the brittle transition in Carrara marble. Journal of Geophysical Research, 94, 4129–4145.CrossRef
Zurück zum Zitat Hadamard, J. S. (1903). Lecons sur la Propagation des Ondes et les Equations de l’Hydrodynamique - Chapitre 6 (241–262 pp.). Librairie Scientifique A. Hermann, Paris.
Zurück zum Zitat Handley, M. F., de Lange, J. A. J., Essrich, F., & Banning, J. A. (2000). A review of the sequential grid mining method employed at Elandsrand Gold Mine. The Journal of The Southern African Institute of Mining and Metallurgy, 100(3), 157–168.
Zurück zum Zitat Hill, R. (1958). A general theory of uniqueness and stability in elastic-plastic solids. Journal of the Mechanics and Physics of Solids, 6, 236–249.CrossRef
Zurück zum Zitat Hirata, T., Satoh, T., & Ito, K. (1987). Fractal structure of spatial distribution of microfracturing in rock. Geophysical Journal of the Royal Astronomical Society, 90(2), 369–374.CrossRef
Zurück zum Zitat Hobbs, B. E., Mulhaus, H.-B., & Ord, A. (1990). Instability, softening and localization of deformation. In R. J. Knipe, & E. H. Rutter (Eds.), Deformation mechanisms, rheology and tectonics (vol. 54, pp. 143–165). Geological Society, Special Publication.CrossRef
Zurück zum Zitat Jessell, M. W., & Lister, G. S. (1991). Strain localisation behaviour in experimental shear zones. Pure and Applied Geophysics, 137(4), 421–438.CrossRef
Zurück zum Zitat Kaneko, K., & Tsuda, I. (2001). Complex systems: Chaos and beyond - A constructive approach with applications in life sciences (273 p.). Springer.
Zurück zum Zitat Kato, N., Yamamoto, K., Yamamoto, H., & Hirasawa, T. (1992). Strain rate effect on frictional strength and the slip nucleation process. Tectonophysics, 211, 269–282.CrossRef
Zurück zum Zitat Keilis-Borok, V. I. (1994). Symptoms of instability in a system of earthquake-prone faults. Physica D, 77, 193–199.CrossRef
Zurück zum Zitat Kostrov, B. V. (1974). Seismic moment and energy of earthquakes and seismic flow of rock. Izvestiya, Physics of the Solid Earth, 13, 13–21.
Zurück zum Zitat Kostrov, B. V., & Das, S. (1988). Principles of earthquake source mechanics. Cambridge University Press.
Zurück zum Zitat Labuz, J. F., Shah, S. P., & Dowding, C. H. (1985). Experimental analysis of crack propagation in granite. International Journal of Rock Mechanics and Mining Sciences, 22(2), 85–98.CrossRef
Zurück zum Zitat Main, I. G., Meredith, P. G., & Sammonds, P. (1992). Temporal variations in seismic event rate and b-values from stress corrosion constitutive laws. Tectonophysics, 211(1-4), 233–246.CrossRef
Zurück zum Zitat Mendecki, A. J. (1993). Real time quantitative seismology in mines: Keynote Address. In R. P. Young (Ed.), Proceedings 3rd International Symposium on Rockbursts and Seismicity in Mines, Kingston, Ontario, Canada (pp. 287–295). Balkema, Rotterdam.
Zurück zum Zitat Mendecki, A. J. (1997). Quantitative seismology and rock mass stability. In A. J. Mendecki (Ed.), Seismic monitoring in mines (1st edn., chap. 10, pp. 178–219). Chapman and Hall.
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 Muhlhaus, H. B., Hobbs, B. E., & Ord, A. (1992). Evolution of fractal geometries in deforming material. In J. R. Tillerson, & W. R. Wawersik (Eds.), Rock Mechanics (pp. 681–690). Balkema, Rotterdam.
Zurück zum Zitat Nicolis, G., & Prigogine, I. (1977). Self-organization in nonequilibrium systems. From dissipative structures to order through fluctuations (491 p.). John Wiley and Sons.
Zurück zum Zitat Nicolis, G., & Prigogine, I. (1989). Exploring complexity. W. H. Freeman and Company.
Zurück zum Zitat Ohnaka, M. (1992). Earthquake source nucleation: A physical model for short-term precursors. Tectonophysics, 211(1-4), 149–178.CrossRef
Zurück zum Zitat Olson, W. A. (1992). The formation of the yield surface vertex in rock. In H. R. Tillerson, & W. A. Wawersik (Eds.) Rock mechanics (pp. 701–705). Balkema, Rotterdam.
Zurück zum Zitat Pearson, G. E. (1956). General theory of elastic stability. Quarterly of Applied Mathematics, 14(2), 133–144.CrossRef
Zurück zum Zitat Prigogine, I. (1980). From eeing to becoming. Time and complexity in the physical sciences. W. H. Freeman and Company.
Zurück zum Zitat Prigogine, I. (1996). Time, chaos and the laws of Nature. In P. Weingartner, & G. Schurz (Eds.), Law and prediction in the light of chaos research, Lecture Notes in Physics (vol. 473, pp. 3–9). Springer.
Zurück zum Zitat Prigogine, I. (1997). The end of certainty: Time, chaos and the laws of nature. The Free Press.
Zurück zum Zitat Rebuli, D., & van Aswegen, G. (2013). Sort term seismic hazard assessment in s.a. gold mines. In A. Malovichko, & D. A. Malovichko (Eds.), Proceedings 8th International Symposium on Rockbursts and Seismicity in Mines, St Petersburg-Moscow, Russia. ISBN:978-5-903258-28-4.
Zurück zum Zitat Reiner, M. (1964). The Deborah number. Physics Today, 62.
Zurück zum Zitat Rudnicki, J., & Rice, J. (1975). Conditions for the localization of deformation in pressure sensitive dilatant materials. Journal of the Mechanics and Physics of Solids, 23, 371–394.CrossRef
Zurück zum Zitat Rudnicki, J. W. (1977). The inception of faulting in a rock mass with a weakened zone. Journal of Geophysical Research, 82, 844–854.CrossRef
Zurück zum Zitat Rudnicki, J. W. (1988). Physical models of earthquake instability and precursory processes. Pure and Applied Geophysics, 126(2-4), 531–554.CrossRef
Zurück zum Zitat Scholz, C. H. (1990). The mechanics of earthquakes and faulting. Cambridge University Press.
Zurück zum Zitat Stuart, W. D. (1981). Stiffness method for anticipating earthquakes. Bulletin of the Seismological Society of America, 71(1), 363–370.
Zurück zum Zitat Truesdell, C., & Noll, W. (2004). The non-linear field theories of mechanics (3rd edn., 602 p.). Springer-Verlag.
Zurück zum Zitat van Aswegen, G. (2005). Routine seismic hazard assessment in some South African mines. In Y. Potvin, & M. Hudyma (Eds.), Proceedings 6th International Symposium on Rockbursts and Seismicity in Mines, Perth, Australia.
Zurück zum Zitat van Aswegen, G., & Butler, A. G. (1993). Applications of quantitative seismology in South African gold mines. In R. P. Young (Ed.), Proceedings 3rd International Symposium on Rockbursts and Seismicity in Mines, Kingston, Ontario, Canada (pp. 261–266). Balkema, Rotterdam. ISBN: 90-5410320-5.
Zurück zum Zitat van Aswegen, G., & Mendecki, A. J. (1999). Mine layout, geological features and seismic hazard. Final report gap 303, Safety in Mines Research Advisory Committee, South Africa (pp. 1–91).
Zurück zum Zitat van Aswegen, G., Mendecki, A. J., & Funk, C. (1997). Application of quantitative seismology in mines. In A. J. Mendecki (Ed.), Seismic monitoring in mines (1st edn., chap. 11, pp. 220–245). Chapman and Hall.
Zurück zum Zitat Wadell, H. (1933). Sphericity and roundness of rock particles. The Journal of Geology, 41(3), 310–331.CrossRef
Zurück zum Zitat Wallerstein, I. (1999). The end of the world as we know it: social science for the twenty first century. University of Minnesota Press.
Zurück zum Zitat Wawersik, W. R. J. W. R., Olsson, W. A., Holcomb, D. J., & Chau, K. T. (1990). Localization of deformation in brittle rock: Theoretical and laboratory investigations. In S. P. Shah (Ed.), Proceedings of the International Conference on Micromechanics and Failure. Elsevier.
Zurück zum Zitat Wiener, P. P. (1951). Leibniz selections. Charles Scribner’s Sons.
Zurück zum Zitat Zoller, G., Hainz, S., & Kurths, J. (2001). Observation of growing correlation length as an indicator for critical point behavior prior to large earthquakes. Journal of Geophysical Research, 106(B2), 2167–2175.CrossRef