1: \documentclass[aps]{revtex4}
2: %\documentclass[11pt]{article}
3: \usepackage{graphicx}
4: \setlength{\topmargin}{-1cm}
5: \setlength{\headsep}{2.2cm}
6: \setlength{\evensidemargin}{0cm}
7: \setlength{\oddsidemargin}{0cm}
8: \setlength{\textheight}{23cm}
9: \setlength{\textwidth}{16cm}
10:
11: \ProvidesPackage{times}
12: [1999/03/29 PSNFSS v.7.2
13: Times font as default roman
14: : S Rahtz]
15: \renewcommand{\sfdefault}{phv}
16: \renewcommand{\rmdefault}{ptm}
17: \renewcommand{\ttdefault}{pcr}
18:
19: \begin{document}
20: \renewcommand{\thefootnote}{\fnsymbol{footnote}}
21: \sloppy
22: \newcommand{\rp}{\right)}
23: \newcommand{\lp}{\left(}
24: \newcommand \be {\begin{equation}}
25: \newcommand \ba {\begin{eqnarray}}
26: \newcommand \ee {\end{equation}}
27: \newcommand \ea {\end{eqnarray}}
28:
29: \title{Diffusion of epicenters of earthquake aftershock, Omori law
30: and generalized continuous-time random walk models}
31: \thispagestyle{empty}
32:
33: \author{A. Helmstetter$^1$ and D. Sornette$^{2,3}$\\
34: $^1$ Laboratoire de G{\'e}ophysique Interne et Tectonophysique,
35: Observatoire de Grenoble, Universit{\'e} Joseph Fourier, BP 53X, 38041
36: Grenoble Cedex, France \\ e-mail: ahelmste@obs.ujf-grenoble.fr\\
37: $^2$ Laboratoire de Physique de la Mati\`{e}re Condens\'{e}e\\ CNRS UMR6622 and
38: Universit\'{e} de Nice-Sophia Antipolis\\ B.P. 71, Parc
39: Valrose, 06108 Nice Cedex 2, France\\
40: $^3$ Institute of Geophysics and
41: Planetary Physics and Department of Earth and Space Science\\
42: University of California, Los Angeles, California 90095\\
43: email: sornette@ess.ucla.edu}
44:
45: \begin{abstract}
46:
47: The epidemic-type aftershock
48: sequence model (ETAS) is a simple stochastic process modeling
49: seismicity, based on the two best-established empirical laws, the
50: Omori law (power law decay $\sim 1/t^{1+\theta}$ of seismicity after
51: an earthquake)
52: and Gutenberg-Richter law (power law distribution of earthquake energies).
53: In order to describe also the space distribution of seismicity, we use
54: in addition a power law distribution $\sim 1/r^{1+\mu}$ of distances between
55: triggered and triggering earthquakes.
56: The ETAS model has been studied for the last two decades to model
57: real seismicity catalogs and to obtain short-term probabilistic forecasts.
58: Here, we present an exact mapping between the ETAS model and a class
59: of CTRW (continuous time random walk) models, based on the identification
60: of their corresponding Master equations.
61: This mapping allows us to use the wealth of results previously
62: obtained on anomalous diffusion of CTRW. After translating into the relevant
63: variable for the ETAS model, we provide a classification of the different
64: regimes of diffusion of seismic activity triggered by a mainshock.
65: Specifically, we derive the relation between the average distance between
66: aftershocks and the mainshock as a function of the time from the mainshock
67: and of the joint probability distribution of the times and locations of
68: the aftershocks.
69: The different regimes are fully characterized by the two exponents
70: $\theta$ and $\mu$.
71: Our predictions are checked by careful numerical simulations.
72: We stress the distinction between the ``bare'' Omori law describing the
73: seismic rate activated directly by a mainshock and the ``renormalized''
74: Omori law taking into account all possible cascades from mainshocks
75: to aftershocks
76: of aftershock of aftershock, and so on. In particular, we predict that
77: seismic diffusion or sub-diffusion occurs and should be observable only
78: when the observed Omori exponent is less than $1$, because this signals
79: the operation of the renormalization of the bare Omori law, also at the
80: origin of seismic diffusion in the ETAS model.
81: We present new predictions and insights provided by the ETAS to CTRW mapping
82: that suggest novel ways for studying seismic catalogs. Finally, we
83: discuss the present
84: evidence for our predicted sub-diffusion of seismicity triggered by a
85: main shock,
86: stressing the caveats and limitations of previous empirical works.
87:
88: \end{abstract}
89:
90: \maketitle
91:
92:
93: \pagenumbering{arabic}
94:
95: \section{Introduction}
96:
97: The spatio-temporal complexity of earthquakes is often invoked as
98: an illustration of the phenomenon of critical self-organization with
99: scale-invariant properties
100: \cite{Geilo,Rundleklein,Main,Sorreview,Turcottereview}.
101: This concept points to the importance of developing a system approach in which
102: large scale properties can emerge from the repeating interactions occurring
103: at smaller scales. Such ideas are implemented
104: in models proposing links between the physics of earthquakes
105: and concepts of statistical physics, such as
106: critical points, self-organized criticality, spinodal decomposition,
107: critical depinning, etc., in order to explain the most
108: solidly established facts in
109: the phenomenology of earthquakes, of which we cite the three most important.
110: \begin{itemize}
111: \item LAW 1: The Gutenberg-Richer law \cite{GR} states that the cumulative
112: distribution of earthquake magnitudes $m$
113: sampled over broad regions and large time intervals is proportional
114: to $10^{-b m}$, with a $b$-value $b \approx 1$. Translating into energies $E$
115: with the correspondence
116: $m=(2/3) \log_{10} E + $ constant leads to a
117: power law $\sim 1/E^{B}$ with $B \approx 2/3$.
118:
119: \item LAW 2: Omori law for aftershocks \cite{Omori} states that the
120: rate of earthquakes
121: triggered by a mainshock decays with time according to an inverse power
122: $1/t^{p}$ of time with an exponent $p \approx 1$.
123: \item LAW 3: The earthquakes are clustered in space along hierarchical fault
124: structures \cite{Ouillonetal} and their spatial distribution over
125: long times can be approximately
126: described by a fractal dimension close to $2.2$ (in three dimensions)
127: \cite{KK80}.
128: \end{itemize}
129: There are many other empirical ``laws'' but these three characterize the
130: very fundamentals of seismicity in size, time and space.
131:
132: We should immediately point out that these three laws come with
133: significant caveats.
134: \begin{enumerate}
135: \item There have been on-going controversies on the universality
136: of the exponent $B$ or $b$-value of the Gutenberg-Richter law \cite{Pissor,kaganuniv}.
137: % On the other hand, Kagan \cite{kaganuniv} is a
138: % vigorous proponent of the universality
139: % of $B \approx 2/3$ for all zones.
140:
141: \item The exponent $p$ of Omori law exhibits a large variability from one
142: aftershock sequence to another aftershock sequence and
143: is found typically in the range from $0.3$ to $2$.
144: We note however that not all these values, especially the extreme ones,
145: automatically reflect a bona-fide power law decay and one should exert
146: caution in attributing too much confidence to them.
147:
148: \item The view that geological faults and earthquake hypocenters are fractal
149: objects is now recognized to be a naive description of a much more
150: complex reality in which a hierarchy of scales occur with possibly
151: different organizations at different scales \cite{Ouillonetal}.
152: \end{enumerate}
153:
154: In addition, a major difficulty for making progress in modeling and
155: predicting earthquakes
156: is that these three and other laws may be ``explained'' by a large variety of
157: models, with many different mechanisms. For instance,
158: with respect to the first two laws, we observe the
159: following.
160: \begin{itemize}
161: \item There are many mechanisms that create a power
162: law distribution of earthquake sizes (see for instance the list of
163: mechanisms described in chapter 14 of Ref.~\cite{Sorbook}).
164: \item Omori law
165: is essentially a slowly decaying ``propagator'' describing
166: a long-time memory of past events impacting on the
167: future seismic activity. Such slow power law time decay of the
168: Omori propagator
169: may result from several and not necessarily exclusive mechanisms
170: (see \cite{Harrisreview}
171: and references therein): pore-pressure changes due to pore-fluid flows coupled
172: with stress variations, slow redistribution of stress by aseismic creep,
173: rate-and-state dependent friction within faults, coupling between
174: the viscoelastic lower crust and the brittle upper crust, stress-assisted
175: micro-crack corrosion \cite{Yamakno,Leesor}, slow tectonic driving of
176: a hierarchical geometry
177: with avalanche relaxation dynamics \cite{Huangetal}, etc.
178: \end{itemize}
179:
180: The zeroth order description of earthquakes is to consider a single isolated
181: homogeneous fault on which earthquakes are recurrent to accommodate
182: the long-term slow tectonic loading. But faults are not isolated and
183: the most conspicuous observation is that earthquakes interact and
184: influence each other on complex fault structures. Understanding these
185: interactions is essential
186: for understanding earthquakes and fault self-organization. However, the
187: full impact of interactions between earthquakes is still far
188: from being well understood. The simplest and clearest observation of
189: earthquake interaction is provided by aftershocks whose phenomenology
190: is captured by Omori law (LAW 2). Indeed,
191: aftershocks are the most obvious and striking signature of the clustering
192: of the seismicity in time and space, and are observed after all
193: large shallow earthquakes.
194: Most aftershocks are triggered a few hours or days after the mainshock.
195: However, due to the very slow power law decay of the rate of aftershock,
196: known as the Omori law \cite{Omori}, aftershocks can be triggered
197: up to a hundred years after the mainshock \cite{Utsu}.
198: Aftershocks often occur near the rupture zone of the mainshock
199: with a variety of focal mechanisms suggesting
200: that they are actually on separate structures
201: \cite{BathRichter,Beroza}. They are also sometimes
202: triggered at very large
203: distances from the mainshock \cite{Hill,Steeples,KJ1,Meltzner1,Dreger}.
204: As an example, Hill et al. \cite{Hill} observed aftershocks
205: of the Landers earthquake as far as 1250 km from the epicenter.
206: Similarly to the temporal distribution of aftershocks, a power-law distribution
207: seems to describe well the distribution of distances between pairs
208: of events \cite{KJ1} .
209: Since a power-law decays slowly, it describes a slow decay of the
210: probability of
211: observing aftershocks at large distances to the mainshock.
212:
213: Thus, Omori law can be considered as the simplest and best established
214: description of earthquake interactions of a certain kind. The question we
215: investigate is whether it can be used fruitfully to explain a larger
216: variety of earthquake interactions beyond the class of observations that
217: were used to establish it. In a series of papers \cite{SS,HS1,SH2},
218: we find that
219: Omori law for aftershocks plus the constrain that aftershocks
220: are distributed according to the Gutenberg-Richter power law for
221: earthquake size distribution {\it independently} of the magnitude of their
222: progenitor is enough to derive many of the other empirical ``laws,'' as well
223: the variability of the $p$-exponent.
224: Here, we test the potential of this approach to account for
225: and to quantify observations on aftershock diffusion.
226:
227: Aftershock diffusion refers to the phenomenon of
228: expansion or migration of aftershock zone
229: with time
230: \cite{Mogi,Imoto,Chatelain,Tajima1,Tajima2,Wesson,Ouchi,Noir,Jacques}.
231: Immediately after the
232: mainshock occurrence, most aftershocks are located close to the rupture
233: plane of the mainshock, then aftershocks seem to migrate away from
234: the mainshock,
235: at velocity ranging from 1 km/hour to 1 km/year \cite{Jacques,Rydelek}.
236: Note that this expansion is not universally observed,
237: but is more important in some areas than in others \cite{Tajima1}.
238:
239: The diffusion of aftershocks is usually interpreted as a diffusion of
240: the stress
241: induced by the mainshock, either by a viscous relaxation process
242: \cite{Rydelek},
243: or due to fluid transfer in the crust \cite{Nur,Hudnut,Noir}.
244: Another interpretation of the expansion of aftershocks is given by
245: Dieterich \cite{Diete},
246: who reproduces the Omori law decay of aftershocks and the expansion of the
247: aftershock zone with time, using a rate and state friction law and
248: assuming that the
249: rate of aftershocks is proportional to the stress rate.
250: In his model, the expansion of aftershock zone arises from the
251: non-uniform stress
252: induced by the mainshock.
253: Another alternative explanation is
254: that the diffusion of aftershocks is mainly due to the occurrence
255: of large aftershocks, and to the localization of secondary
256: aftershock close to
257: the largest aftershocks, as observed by Ouchi \cite{Ouchi}.
258: The apparent diffusion of the seismicity may thus result from a
259: cascade process;
260: the mainshock triggers aftershocks that in turn trigger their own
261: aftershocks, and thus lead to an expansion of the aftershock zone.
262:
263: In the present paper, we investigate the epidemic time aftershock
264: sequence model (ETAS),
265: and show that the cascade of secondary aftershocks can indeed explain
266: the reported diffusion of aftershocks.
267: The ETAS model was introduced by Kagan and Knopoff \cite{KK}
268: (in a slightly different form than used here) and Ogata \cite{Ogata1}
269: to describe the temporal and spatial clustering of seismicity.
270: This model provides a tool for understanding the
271: clustering of the seismic activity, without arbitrary distinction between
272: aftershocks, foreshocks and mainshocks.
273: In this model, all earthquakes are assumed to be simultaneously
274: mainshocks, aftershocks and possibly foreshocks. Each earthquake generates
275: aftershocks that decay with time according to Omori law, which
276: will in turn generate their own aftershocks. The seismicity rate at
277: any given time and location is given by
278: the superposition of aftershock sequences of all events impacting that region
279: at that time according to space-time ``propagators.''
280: The additional ingredient in the version of the ETAS model that we
281: study is that the number of aftershocks per earthquake increases exponentially
282: $\propto 10^{\alpha m}$ with the magnitude $m$ of
283: the mainshock (i.e., as a power law $\propto E^{2 \alpha /3}$
284: of the energy released by the mainshock), in agreement with the observations
285: \cite{YaShi,Drakatos}.
286: Since the energy of an earthquake is a power
287: law of its rupture length, this law expresses
288: the very reasonable idea that the number of events related to a given
289: earthquake is proportional to a power of its volume of influence.
290: The value of the exponent $\alpha$ controls the nature of
291: the seismic activity, that is, the relative role of small compared
292: to large earthquakes. Few studies have measured $\alpha$
293: in seismicity data \cite{YaShi,Guo2,H02}.
294: This parameter $\alpha$ is often found close to $b$ \cite{YaShi}
295: or fixed arbitrarily equal to $b$ \cite{KK,Felzer}.
296: In the case where $\alpha$ is close to the Gutenberg-Richter $b$-value,
297: this law also reproduces
298: \cite{Felzer} the self-similar empirical Bath' s law \cite{Bath},
299: which states that the
300: average difference $m_M-m_A$ in size between a mainshock and its largest aftershock is
301: $1.2$ magnitude units, regardless of the mainshock magnitude:
302: $m_A = m_M - 1.2$.
303: If $\alpha < b$, small earthquakes, taken together, trigger more
304: aftershocks than larger
305: earthquakes. In contrast, large earthquakes dominate earthquake
306: triggering if $\alpha \geq b$.
307: This case $\alpha \geq b$ has been studied analytically in the
308: framework of the ETAS model by Ref. \cite{SH2}
309: and has been shown to eventually lead to a finite time singularity of
310: the seismicity rate. This explosive regime cannot however describe
311: a stationary seismic activity.
312:
313: A natural way to tame this singular behavior is to introduce an
314: upper cut-off for the magnitude distribution at large magnitudes,
315: mirroring the cut-off $m_0$ used for the low-magnitude range. The
316: physical argument for introducing this cut-off is based on the finiteness
317: of the maximum earthquake that the earth is capable of carrying.
318: The specific way of introducing such a cut-off (abrupt or smooth
319: with a transition to a power law with larger exponent or to an exponential taper) is not
320: very important qualitatively because all these laws will regularize the singular
321: behavior and make the average branching ratio finite. Such
322: regularization with a maximum upper magnitude then allows $\alpha \geq b$.
323: The special case $\alpha = b$ required for Bath's law to hold exactly
324: can not therefore be excluded.
325:
326: However, based on a recent re-analysis of seismic catalogs using the powerful
327: collapse technique, one of us \cite{H02} has presented strong evidence
328: that $\alpha$ is strictly smaller than $b$.
329: In this paper, we will therefore consider only the case $\alpha<b$
330: and take $\alpha=0.5$ specifically in our numerical simulations.
331: In this regime $\alpha<b$, Bath' s law cannot be reproduced because
332: the average difference in size between a mainshock and its largest aftershock
333: increases with the mainshock magnitude. For $\alpha < b$, it is
334: easy to show that Bath's law
335: is replaced by $m_A = (\alpha/b) m_M -$ constant, where $m_M$ and
336: $m_A$ are the magnitudes of the mainshock and of the largest aftershock.
337: Tests of this prediction will be reported in a future publication but
338: we expect that distinguishing this
339: modified Bath's law from Bath's law will be a difficult task due to the
340: limited range of the studied magnitudes as well as the dependence
341: of the distribution of $m_M-m_A$ on the magnitude thresholds chosen for the mainshocks and
342: for the aftershocks \cite{consolebath}.
343: %In addition,
344: %the ``pure''~ Bath' s law may be an artefact of the method used to define
345: %mainshocks and aftershocks \cite{VJ69}.
346:
347: We assume that the distribution of all earthquakes follow the
348: Gutenberg-Richter distribution and take this
349: distribution of aftershock sizes to be independent of the magnitude
350: of the mainshock. Therefore,
351: an earthquake can trigger a larger earthquake, albeit with
352: a small probability. This model can thus describe a priori
353: both aftershock and foreshock sequences.
354: The ETAS model has been calibrated to real seismicity
355: catalogs to retrieve its parameters
356: \cite{Ogata1,Ogata2,Ogata3,Ogata5,Ogata6,Kagan,Guo2,Felzer}
357: and to give short-term probabilistic forecast of
358: seismic activity by extrapolating past seismicity into the future
359: via the use of its space-time propagator \cite{KK,KJ2,Console}.
360:
361: The ETAS model is a branching model which exhibits different
362: regimes \cite{HS1} depending upon the value of
363: the branching ratio $n$, defined by the
364: average number of primary aftershocks per earthquake.
365: The critical case $n=1$ corresponds to exactly one primary aftershock
366: per earthquake, when averaging over all mainshock magnitudes larger
367: than a threshold $m_0$. Let us stress that $n$ is an average quantity
368: which does not reflect adequately the large variability of the number of
369: aftershocks per main shock, as a function of its magnitude. Indeed,
370: the number of aftershocks per mainshock
371: increases exponentially fast as a function of the mainshock magnitude,
372: so that large mainshocks
373: will have significantly more than $n$ aftershocks. For $\alpha =0.5$,
374: a magnitude $7$-earthquake gives typically $10$
375: times more direct aftershocks than a magnitude $5$, and $100$ times more
376: direct aftershocks than a magnitude $3$-earthquake. The increase
377: in triggered seismic activity with the magnitude of the mainshock
378: is obviously stronger for a larger value of $\alpha$. Note that these
379: numbers refer to aftershocks of the first generation; the total number
380: of triggered events is larger by the factor $1/(1-n) \sim 10$
381: (for $n \approx 0.9$ which is typical),
382: due to the cascades of secondary aftershocks.
383: Notwithstanding this large variability, the average
384: number $n$ of primary aftershocks per earthquake controls the global regime.
385: For $n$ exactly equal to $1$,
386: seismicity is at the border between death and growth.
387: In the sub-critical regime $n<1$, since each earthquake triggers on
388: average less
389: that one aftershock, starting from a large event, the seismicity will decrease
390: with time and finally die out.
391: The super-critical $n>1$ corresponds to more that one primary aftershock per
392: earthquake on average. Starting from a large earthquake, after a transient
393: regime, the average seismicity will finally increase exponentially
394: with time \cite{HS1},
395: but there is still a finite probability for aftershock sequences to die out.
396:
397: The numerical simulations reported below have been performed with $\alpha=0.5$.
398: It is probable that a good fit to seismic
399: data is obtained by using a value of $\alpha \approx 0.8$ larger that the value $0.5$,
400: as reviewed and documented recently by one of us \cite{H02}.
401: We have checked that results similar to those presented below hold true qualitatively for larger
402: values $0.5< \alpha <1$. Such larger values of $\alpha$ lead however to stronger fluctuations
403: which are more difficult to handle numerically because the variance of
404: the number $\rho(m)$ of direct triggered aftershocks defined below in
405: (\ref{formrho}) becomes undefined for $\alpha > 0.5$.
406: A full understanding of this regime requires a special treatment
407: that will be reported elsewhere.
408:
409: Sornette and Sornette \cite{SS} studied analytically a particular case of this
410: model, without magnitude and spatial dependence, and they considered only the
411: subcritical regime $n<1$. Starting with one event at time $t=0$ and
412: considering that each earthquake generates an aftershock sequence
413: with a ``local'' Omori exponent $p=1+\theta$, where $\theta>0$,
414: they studied the decay law of the ``global'' aftershock sequence,
415: composed of all secondary aftershock sequences, i.e., by taking into
416: account that the primary aftershocks can create secondary aftershocks
417: which themselves may trigger tertiary aftershocks and so on.
418: They found that the global aftershock rate decays according to an Omori law
419: with an exponent $p=1-\theta <1$,
420: up to a characteristic time \cite{SS,HS1}
421: \be
422: t^* = c ~\left({n~\Gamma(1-\theta) \over |1-n|}\right)^{1/\theta}~,
423: \label{tstar}
424: \ee
425: and then recovers the local Omori exponent
426: $p=1+\theta$ for time larger than $t^*$.
427: Helmstetter and Sornette \cite{HS1} extended their analysis to the general ETAS
428: model with magnitude dependence, and considered both the sub- and the
429: super-critical regime, but still restricted the analysis to the temporal
430: distribution of the seismicity, without spatial dependence.
431: In the sub-critical regime, they recovered
432: the crossover found by Sornette and Sornette \cite{SS}.
433: In addition, Helmstetter and Sornette \cite{HS1} give the explicit mathematical
434: formula for the gradual transition between the Omori law with exponent
435: $p=1-\theta$ for $t \ll t^*$ to the Omori law with exponent
436: $p=1+\theta$ for $t \gg t^*$. This smooth transition can be observed
437: in figure \ref{n} on the line calculated for $t^*=10^9$ days
438: with $n<1$. $t^*$ can thus be viewed as the time where the
439: apparent exponent $p$ of the Omori law is
440: approximately in between the two asymptotic values $1-\theta$
441: and $1+\theta$. A more rigorous
442: mathematical definition \cite{HS1} is that $t^*$ is the
443: characteristic time scale such that $\beta t^*$ is the dimensionless
444: variable of the Laplace transform (with variable $\beta$) of the seismicity rate.
445:
446: In the super-critical
447: regime, Helmstetter and Sornette \cite{HS1}
448: found a novel transition between a power-law decay with exponent
449: $p=1-\theta$ at early times, similar to the sub-critical regime, to
450: an exponential increase of the seismicity at large times.
451: The regime where $\alpha > b$ or equivalently $2 \alpha/3 > B$ has been found
452: to lead to a new kind of critical stochastic finite-time-singularity
453: \cite{SH2}, relying on the interplay
454: between long-memory and extreme fluctuations. Recall that
455: the number of aftershocks per earthquake increases as a power law
456: $\propto E^{2 \alpha/3}$ of the energy released by the mainshock
457: whereas the number of earthquakes of energy $E$ decreases as
458: the Gutenberg-Richter law $\propto 1/E^{1+B}$.
459: Intuitively, when $2 \alpha /3 > B$,
460: the increase in the rate of creation of aftershocks with the mainshock
461: energy more than compensate the decrease of the probability to get
462: a large mainshock when the mainshock energy increases.
463: This theory based solely on the ETAS model has been found to account for the
464: main observations (power law acceleration and discrete scale invariant
465: structure) of critical rupture of heterogeneous materials, of the largest
466: sequence of starquakes ever attributed to a neutron star as well as of some
467: earthquake sequences \cite{SH2}.
468:
469: In the sequel, we extend the analytical study of the temporal ETAS
470: model \cite{SS,HS1,SH2}
471: to the spatio-temporal domain. To model the spatial distribution of
472: aftershocks, we assume that the distance between a mainshock and each
473: of its direct aftershock is drawn from a given distribution,
474: independently of the magnitude of the mainshock and of
475: the delay between the mainshock and its aftershocks. For illustration but
476: without loss of generality for the mapping to the
477: continuous time random walks model (CTRW) discussed later, we
478: shall take a power law distribution of distances between earthquakes.
479: We take the simplest and most parsimonious hypothesis that
480: space, time and magnitude are decoupled in the earthquake propagator.
481: Our first result is to establish a correspondence between
482: the ETAS model and the CTRW, first introduced
483: by Montroll and Weiss \cite{Montroll1} and used to model many
484: physical processes.
485: We then build on this analogy to derive the joint
486: probability distribution of the times and locations of aftershocks.
487: We show analytically that, for sufficiently short times $t<t^*$,
488: the average distance between a mainshock and its aftershock increases
489: subdiffusively
490: as $R \sim t^H$, where the exponent $H$ depends on the local Omori
491: exponent $1+\theta$ and
492: on the distribution of the distances between an earthquake and its aftershocks.
493: We also demonstrate that the local Omori law is not universal, but varies as
494: a function of the distance from the mainshock. Due to the diffusion
495: of aftershocks with time, the decay of aftershock is faster close
496: to the mainshock than at large distances. These non-trivial space-time
497: couplings occur notwithstanding the decoupling
498: between space, time and magnitude in the ``bare'' propagator, and is due to
499: the existence of cascades of aftershocks.
500:
501: A recent work of Krishnamurthy et al. \cite{Tanguy} substantiates the general
502: modeling strategy used here of representing the space-time dynamics of earthquakes
503: by an effective stochastic process (the ETAS model) entirely defined
504: by two exponents
505: (corresponding to our $\mu$ and $H(\theta, \mu)$ defined below), where $\mu$ is
506: the exponent of
507: the power law distribution of jumps between successive active sites and $H$
508: is the (sub-)diffusion exponent.
509: Indeed, Krishnamurthy et al. \cite{Tanguy} show
510: that the Bak and Sneppen model and the Sneppen model of extremal dynamics
511: (corresponding to a certain class of self-organized critical behavior
512: \cite{Sorbook})
513: can be completely characterized by a suitable stochastic process called
514: ``Linear fractional stable motion.'' Beyond recovering the scaling exponents
515: of this model, the stochastic process strategy predicts the conditional
516: probabilities of successive activations at different sites and thus offers
517: novel insights. We note that this approach with the
518: Linear fractional stable motion is extremely close in spirit as well as
519: in form to our approach mapping the ETAS model to the CTRW model.
520: The ETAS model can thus be taken to represent an effective stochastic process
521: of the complex self-organization of seismicity.
522:
523: \section{The ETAS model}
524:
525: \subsection{Definitions and specific parameterization of the ETAS model}
526:
527: We assume that a given event (the ``mother'') of magnitude
528: $m_i$ occurring at time $t_i$ and position $\vec r_i$
529: gives birth to other events (``daughters'') of any possible
530: magnitude chosen with some independent Gutenberg-Richter
531: distribution at a later time
532: between $t$ and $t+dt$ and at point $\vec r \pm \vec dr$ to within
533: $d\vec r $ at the rate
534: \be
535: \phi_{m_i}(t-t_i, \vec r-\vec r_i) = \rho(m_i)~\Psi(t-t_i)~\Phi(\vec
536: r-\vec r_i)~.
537: \label{first}
538: \ee
539: We will refer to $\phi_{m_i}(t-t_i, \vec r-\vec r_i)$ both as
540: the seismic rate induced by a single mother or as the ``bare propagator''.
541: It is the product of three independent contributions:
542: \begin{enumerate}
543: \item $\rho(m_i)$ gives the number of daughters born from a mother with
544: magnitude $m_i$. This term will in general be chosen to account for the
545: fact that large earthquakes have many more triggered events than small
546: earthquakes.
547: Specifically, we take
548: \be
549: \rho(m_i) = K ~10^{\alpha (m_i-m_0)}~, \label{formrho}
550: \ee
551: which, as we said earlier, is justified by the power law dependence of the
552: volume of stress perturbation as a function of the earthquake size.
553: $\alpha$ quantifies how fast the
554: average number of daughters per mother increases with the magnitude of the mother.
555:
556: \item $\Psi(t-t_i)$ is a normalized waiting time distribution giving the rate
557: of daughters born at time $t-t_i$ after the mother. The normalization condition
558: reads $\int_0^{+\infty} dt ~\Psi(t) =1$. $\Psi(t-t_i) dt$ can thus
559: be interpreted as the probability for a daughter to be born between $t$ and
560: $t+dt$ from the mother that was born at time $t_i$. $\Psi(t-t_i)$ embodies
561: Omori law: it is the ``bare'' or ``direct'' Omori law
562: \be
563: \Psi(t) = {\theta~c^{\theta} \over (t+c)^{1+\theta}}~H(t)~,
564: \label{psidef}
565: \ee
566: where $\theta>0$ and $H(t)$ is the Heaviside function.
567:
568: \item $\Phi(\vec r-\vec r_i)$ is a normalized spatial ``jump'' distribution from
569: the mother to each of her daughter, quantifying the probability for a daughter
570: to be triggered at a distance $|\vec r-\vec r_i|$ from the mother.
571: Specifically, we take
572: \be
573: \Phi(\vec r) = {\mu \over d ({|\vec r| \over d}+1)^{1+\mu}}~,
574: \label{phidef}
575: \ee
576: which has the form of an (isotropic) elastic Green function
577: dependence describing the stress transfer in an elastic upper crust.
578: The exponent $\mu$ is left adjustable to account for heterogeneity
579: and the possible complex modes of stress transfers.
580: The normalization condition reads $\int d\vec r ~\Phi(\vec r) =1$ where
581: the integral is carried out over the whole space.
582: \end{enumerate}
583:
584: The physical justification for this decoupled model (\ref{first}) in which
585: $\phi_{m_i}(t-t_i, \vec r-\vec r_i)$ is the product of three independent
586: distributions is that elastic waves propagate at kilometers per second
587: and thus almost instantaneously reset the stress field after a large
588: main shock.
589: In other words, there is a well-defined separation of time scales
590: between the time
591: of propagation of seismic waves (seconds to minutes)
592: which control the convergence to a new mechanical
593: equilibrium after the main shock and the time scales involved in
594: aftershock sequences
595: (hours, days, months to many years).
596: The spatial dependence in (\ref{first}) reflects the stress redistribution.
597: This new stress field then relaxes slowly and more or less
598: independently from point to
599: point leading to the local Omori law $\Psi(t-t_i)$. Notwithstanding this
600: argument, the decoupling in (\ref{first})
601: between the local responses in magnitudes, space and time
602: is mostly performed because of its simplicity. It constitutes
603: an approximation that should be checked and relaxed in future
604: studies.
605:
606: We assume a distribution $P(m)$ of earthquake
607: sizes expressed in magnitudes $m$ which follows the Gutenberg-Richter
608: distribution
609: \be
610: P(m) = b~ \ln(10) ~ 10^{-b (m-m_0)}~, \label{gojfwo}
611: \ee
612: with a $b$-value usually close to $1$.
613: $m_0$ is a lower bound magnitude below which no daughter is triggered.
614:
615:
616: \subsection{The branching ratio $n$}
617:
618: A key parameter of the ETAS is the average number $n$ of
619: daughter-earthquakes
620: created per mother-event, summed over all possible magnitudes. As we shall see,
621: it is also natural to call it the ``branching ratio''.
622: To see this,
623: consider the integral of the seismic rate $\phi_{m_i}(t-t_i, \vec r-\vec r_i)$
624: induced by one earthquake
625: over all times after $t_i$, over all spatial positions and over all
626: magnitudes $m_i \geq m_0$,
627: which must give by definition the average number $n$ of direct
628: (or primary) daughter-earthquakes
629: created per mother-event independently of its magnitude.
630: For $\alpha<b$ and using (\ref{first}), (\ref{formrho}) and (\ref{gojfwo}),
631: it is exactly given by
632: \be
633: n \equiv \int d\vec r ~ \int_{t_i}^{+\infty} dt ~\int_{m_0}^{+\infty}
634: dm_i~P(m_i)~
635: \phi_{m_i}(t-t_i, \vec r-\vec r_i) =
636: \int_{m_0}^{+\infty} dm_i~P(m_i)~ \rho(m_i)={K~b\over b-\alpha} ~,
637: \label{second}
638: \ee
639: since the two integrals over time and space contribute
640: each a factor $1$ by the normalization of $\Psi$ and $\Phi$.
641: This result (\ref{second}) is identical to that found
642: in absence of spatial dependence of
643: $\phi_{m_i}(t-t_i)$ with respect to $\vec r-\vec r_i$ due to the
644: factorization of the rate $\rho$,
645: time $\Psi$ and space $\Phi$ dependences \cite{HS1}.
646: The branching ratio has also been evaluated in the case
647: where the magnitude distribution follows a gamma distribution \cite{Kagan}.
648:
649: We stress again that $n$ is an {\it average} quantity which does not reflect
650: the large fluctuations in the number of aftershocks from events to events.
651: Indeed, large events with magnitudes $M$ produce
652: in general many more aftershocks than small events with magnitude $m<M$, simply
653: because $\rho(M) \gg \rho(m)$ if $M>m$ (see the exponential
654: dependence (\ref{formrho}) of $\rho(m)$ on the magnitude $m$).
655:
656: \subsection{Numerical simulation of the spatial ETAS model}
657: \label{num}
658: The ETAS model has been simulated numerically using the algorithm
659: described in Refs.~\cite{Ogata4,Ogata5}.
660: Starting with a large event of magnitude $M$ at time $t=0$, events are
661: then simulated sequentially. At any given time $t$, we calculate the
662: conditional seismic rate $\lambda(t)$ defined by
663: \be
664: \lambda(t)=\sum_{t_i\leq t} K~10^{\alpha(m_i-m_0)}
665: {\theta c^\theta \over (t-t_i+c)^{1+\theta}}
666: \label{mgmkr}
667: \ee
668: where $K = n (b-\alpha)/b$, and $t_i$ and $m_i$ are the
669: times and magnitudes of all preceding events that occurred at time $t_i\leq t$.
670: Note that we use the bare propagator because the sum in (\ref{mgmkr}) is
671: performed exhaustively on the complete catalog of past events.
672: The time of the following event is then determined according to the
673: non-stationary Poisson process of conditional intensity $\lambda(t)$,
674: and its magnitude is chosen according to the Gutenberg-Richter
675: distribution with parameter $b$.
676: To determine the position in space of this new event, we first choose
677: its mother randomly among all preceding events with a probability proportional to
678: their rate of aftershocks $\phi_{m_i}(t-t_i)$ evaluated at the time
679: of the new event. Once the mother has been chosen, we generate the
680: distance $r$ between the new earthquake and its mother according to
681: the power-law distribution $\Phi(\vec r)$ given by (\ref{phidef}).
682: The location of the new event is determined by assuming an isotropic
683: distribution of aftershocks.
684: By this rule, it is clear that new events tend to be close in general
685: to the last large earthquakes, leading to space clustering.
686:
687: Note that this two-steps procedure is equivalent to
688: but more convenient for a numerical implementation than the one-step method,
689: consisting of calculating at each point on a fine
690: space-covering grid the seismic rate, equal to the sum over all
691: preceding mothers
692: weighted by the bare space $\Phi(\vec r)$ and time $\Psi(t)$ propagators
693: given by (\ref{phidef}) and (\ref{psidef}); after normalizing, these rates then
694: provide to each grid point a probability for the event to occur on that point.
695: The equivalence between our two-step procedure and the direct calculation
696: of the seismic rates is based on the law of conditional probabilities:
697: probability of next event ($A$) = probability of next event conditioned on its mother
698: (event $B$) $\times$ probability of choosing the mother, i.e., $P(A,B) = P(A | B)
699: \times P(B)$.
700:
701: Figure \ref{map} shows the result of a numerical simulation of the ETAS model
702: which exhibits a diffusion of the seismic activity.
703: We simulate a sequence of aftershocks and secondary aftershocks starting
704: from a mainshock of magnitude $M=7$, with the following parameters:
705: $\theta=0.2$, $b=1$, $\alpha=0.5$, $n=1$ and $\mu=1$.
706: At early times, aftershocks are localized close to the mainshock,
707: and then diffuse and cluster close to the largest aftershocks. This
708: (sub-)diffusion is extremely slow, as we shall quantify in the sequel.
709: Our purpose is to provide a theory for this process based on the ETAS
710: model. This theory will be tested by numerical simulations.
711:
712:
713: The different regimes are illustrated by Figure \ref{n} which shows
714: the seismicity rate $N(t)$ for the temporal ETAS model studied by
715: \cite{SS,HS1} obtained by summing the seismic activity over all space
716: for the 3 cases $n<1$ (sub-critical), $n=1$ (critical) and $n>1$
717: (super-critical).
718: The sub-critical regime is characterized by the existence of
719: the time scale $t^*$ given by (\ref{tstar}).
720: There is no difference between the critical case $n=1$ and the
721: sub-critical case for $t<t^*$ (see figure \ref{n}). Indeed, the difference
722: between the sub-critical regime and the critical regime can be observed
723: only for $t>t^*$. A simple way to see this is to realize that the critical
724: regime $n=1$ gives $t^* = +\infty$, meaning that, in the critical regime,
725: one is always in the situation $t<t^*$.
726:
727:
728: It is interesting to note that the spatial distribution of epicenters shown in
729: the right panel of figure \ref{map}
730: has the visual appearance of a fractal set of points.
731: This is confirmed by the calculation
732: of the correlation dimension of this set of $N=3000$ points
733: generated in the time interval $[30, 70]$ yrs, which is found
734: approximately equal to $D_2=1.5 \pm 0.05$ over more than
735: two decades in spatial scales, as shown in figure \ref{cordim}.
736: If we use instead all $30,000$ events of the simulation performed
737: up to time $t=70$ yrs, we find
738: $D_2=1.85 \pm 0.05$ while the correlation dimension of the geometrical set
739: made of the epicenters of the $10,000$ last events (time interval $[7, 70]$ yrs) is
740: $D=1.7 \pm 0.05$, also over more than two decades in scale. These values
741: are similar to those reported for 2D maps of active fault systems
742: \cite{ScholzMandel,Anneexp,BartonLapointe1,Ouillonetal},
743: %but cannot be compared directly with the value $d=2.2$ obtained
744: %in Ref.~\cite{KK80} from an analysis of hypocenters.
745: and are in good agreement with $D_2$ values in the range $[1.65, 1.95]$
746: measured for aftershocks epicenters \cite{Nanjo}.
747: The fractal clustering of the earthquake epicenters,
748: according to the ETAS model, occurs because of a self-similar process
749: taking place on many different scales. However, the description
750: of this multi-scale process solely in terms of a single fractal dimension
751: fails to fully embody the complex spatial superposition of
752: local ``singularities'' associated with each aftershock on the one hand
753: and finite-size effects (stemming from the finite lifetime
754: of each aftershock sequence) on the other hand. Each event indeed creates
755: its cloud
756: of direct aftershocks which can be characterized by its singular exponent
757: $1-\mu$ for $\mu \leq 1$ and $0$ for $\mu>1$, defined by the scaling
758: $\propto\int_0^R r dr/ r^{1+\mu} \propto R^{1-\mu}$ of the ``mass''
759: of the cloud with its radius $R$. Finite-size effects and
760: randomness have been documented to generate realistic but
761: sometimes spurious fractal signatures
762: \cite{Ouillonsor,Hamburger,Eneva,Malcai}. This problem requires
763: a special study which is left for another work.
764:
765:
766: \subsection{Relationship with the space-independent ETAS model \label{consosk}}
767:
768: The spatial ETAS model reduces to the space-independent ETAS model solved
769: in \cite{HS1} by integrating the dressed propagator
770: obtained below over all space.
771: In the Fourier representation (see expression (\ref{ngnslqw})), this
772: corresponds to putting the wavenumber $k$ to zero. Indeed,
773: for $k=0$, the Fourier transform amounts to perform a simple integration over
774: all space. Since ${\hat \Phi}(\vec k=\vec 0)=1$, expression (\ref{ngnslqw})
775: derived below reduces to the form studied at length in \cite{HS1}.
776: Therefore, all results reported previously hold also for the version of the
777: space-dependent ETAS model studied here, when averaging over the whole space.
778: This is an important property that all the solutions discussed below must obey.
779:
780:
781: \section{Mapping of the ETAS model on the CTRW model}
782:
783: In order to study the space-time properties of the ETAS model,
784: it is very useful to use an exact correspondence between
785: the ETAS model and the continuous time random walk (CTRW) that we
786: establish here. In this way,
787: we can adapt and use the wealth of results previously derived
788: for the CTRW. But first, let us demonstrate the correspondence
789: between the ETAS and CTRW models. For this, our
790: strategy is to derive the Master equation for both models and show
791: that they are identical.
792:
793: \subsection{The Master equation of the ETAS model}
794:
795: The ETAS model can be rephrased by defining the rate
796: $\phi_{m_i \to m}(t-t_i, \vec r-\vec r_i)$ at which
797: a given event (the ``mother'') of magnitude
798: $m_i \geq m_0$ occurring at time $t_i$ and position $\vec r_i$
799: gives birth to other events (``daughters'') of specified magnitude
800: $m$ at a later time
801: between $t$ and $t+dt$ and at point $\vec r$ to within an
802: infinitesimal volume $|d{\vec r}|$.
803: Note that the only difference with respect to the previous definition
804: (\ref{first})
805: is that we now specify also the magnitude $m$ of the daughter.
806: $\phi_{m_i \to m}(t-t_i, \vec r-\vec r_i)$ is given by
807: \be
808: \phi_{m_i \to m}(t-t_i, \vec r-\vec r_i) = \rho(m_i \to
809: m)~\Psi(t-t_i)~\Phi(\vec r-\vec r_i)~,
810: \label{firstprime}
811: \ee
812: where $\Psi(t-t_i)$ and $\Phi(\vec r-\vec r_i)$ are the same as
813: previously while
814: \be
815: \rho(m_i \to m) = P(m)~ \rho(m_i)~.
816: \ee
817: With the parameterization (\ref{formrho}) and (\ref{gojfwo}), this reads
818: \be
819: \rho(m_i \to m) = n~ \ln(10)~(b - \alpha) ~10^{\alpha
820: (m_i-m_0)}~10^{-b (m-m_0)}~.
821: \label{formrhafdgo}
822: \ee
823:
824: Let us consider the case where there is an
825: origin of time $t=0$ at which we start recording the rate of
826: earthquakes, assuming that
827: a large earthquake has just occurred at $t=0$ and somehow
828: reset the clock. In the
829: following calculation, we will forget about the effect of events
830: at times prior to $t=0$
831: and count all aftershocks that are created only by this main shock.
832:
833: Let us call $N_m(t,\vec r) dt ~dm ~d\vec r$ the number of earthquakes
834: occurring between
835: $t$ and $t+dt$ of magnitude
836: between $m$ and $m+dm$ inside of box of volume $|d{\vec r}|$ centered
837: at point $\vec r$.
838: $N_m(t,\vec r)$ is the
839: solution of a self-consistency equation that formalizes
840: mathematically the following
841: process\,: an earthquake may trigger aftershocks; these aftershocks may trigger
842: their own aftershocks, and so on. The rate of seismicity at a given
843: time $t$ and position $\vec r$ is the
844: result of this cascade process. The self-consistency equation that sums up this
845: cascade reads
846: \be
847: N_m(t,\vec r) = S(t, \vec r, m) +\int {\vec dr}' \int_{m_0}^{\infty} dm'
848: \int_0^t d\tau~ \phi_{m' \to m}(t-\tau,\vec r-{\vec r}') ~N_{m'}(\tau,
849: {\vec r}') ~.
850: \label{third}
851: \ee
852: The rate $N_m(t,\vec r)$ at time $t$ and position $\vec r$ is the sum
853: over all induced rates from all earthquakes of all possible magnitudes
854: that occurred at all previous times and locations propagated to the present
855: time $t$ and to the position $\vec r$ of observation by the corresponding
856: bare propagator. The induced rate
857: of events per earthquake that occurred at an earlier time $\tau$ and
858: position ${\vec r'}$ is equal to $\phi_{m' \to m}(t-\tau,\vec r-{\vec r}')$.
859: The source term $S(t, \vec r)$ is the main shock plus the background
860: seismicity, if any.
861: In absence of background seismicity, a main earthquake which occurs
862: at the origin
863: of time $t=0$ at position $\vec r= \vec 0$ with magnitude $M$ gives
864: \be
865: S(t,\vec r, m)=\delta(t)~\delta(m-M)~\delta(\vec r)
866: \label{S(t,r,m)}
867: \ee
868: where $\delta$ is the Dirac distribution.
869: Other arbitrary source functions can be chosen.
870:
871: The source term corresponding to a single mainshock is
872: indeed the delta function (\ref{S(t,r,m)}) rather than
873: the direct Omori law
874: created by this mainshock in direct lineage. To see this,
875: notice that the direct Omori law
876: is recovered from (\ref{third}) by replacing $~N_{m'}(\tau, {\vec r}')$
877: in the integral by $S(t, \vec r, m)$ given by (\ref{S(t,r,m)}). This
878: shows that the difference between the renormalized
879: and the direct Omori laws comes from taking into account the
880: secondary, tertiary, etc., cascade of aftershocks.
881:
882: As we have seen, a key assumption of the ETAS model is that the
883: daughters born from a
884: given mother have their magnitude drawn independently of the magnitude
885: of the mother and of the process that give them birth, with a
886: probability given by the
887: Gutenberg-Richter distribution (\ref{gojfwo}). The consequences
888: resulting from relaxing this hypothesis will be reported elsewhere.
889: Keeping this assumption, it can be shown \cite{HSG} that for $\alpha \leq b/2$
890: an ensemble of realizations will obey
891: \be
892: N_m(t,\vec r) = P(m) ~N(t,\vec r)~,~~~~{\rm for}~~t>0~, \label{ggnnlalaq}
893: \ee
894: which makes explicit the separation of the magnitude from the time
895: and space variables.
896: $N(t,\vec r)$ is the number of events at position $\vec r$ at time $t$ of any
897: possible magnitude. Expression (\ref{ggnnlalaq}) means that the
898: Gutenberg-Richter
899: distribution is preserved at all times.
900: That (\ref{ggnnlalaq}) holds for the ETAS model stems from
901: the fact that
902: the waiting time $\Psi(t)$ distribution (\ref{psidef}) and jump size
903: $\Phi(\vec r)$
904: distribution (\ref{phidef}) are independent of the magnitudes and that
905: fluctuations in the seismicity rate are not too wild for $\alpha \geq b/2$.
906: Note that, in a more complex model in which time, space and magnitudes are
907: interdependent, expression (\ref{ggnnlalaq}) would become a
908: mean-field approximation,
909: in which the fluctuations of the rates induced by the fluctuations of the
910: realized magnitudes of the daughters factorize from the process.
911:
912: Putting (\ref{ggnnlalaq}) in (\ref{third}), for $t>0$ when the source
913: term $S(t,\vec r, m)$ is identically zero, one can simplify by $P(m)$
914: and obtain
915: \be
916: N(t,\vec r) = \int d\vec r'
917: \int_0^t d\tau~ \phi(t-\tau,\vec r-\vec r') ~N(\tau, \vec r') ~, ~~~~t>0~,
918: \label{thirdterter}
919: \ee
920: where
921: \be
922: \phi(t-\tau,\vec r-\vec r') = \int_{m_0}^{\infty} dm' ~ P(m')
923: \phi_{m'}(t-\tau,\vec r-\vec r')~.
924: \label{fngnflq}
925: \ee
926:
927: Equation (\ref{thirdterter}) is nothing but the expectation
928: (or statistical average, i.e., average over an ensemble of
929: realizations) of expression (\ref{mgmkr}),
930: with the definition $N(t,\vec r) \equiv {\rm E}[\lambda(t)~\Phi(\vec r)]$.
931: Therefore, the Master
932: equation obtained here gives us only the first moment of the space-time
933: dynamics of seismicity. It is not difficult to derive the equations for the
934: variance and covariance of the seismic rate as well as higher moments.
935:
936: The value of the source term at $t=0$ that should be incorporated
937: in (\ref{thirdterter}) requires more care. Indeed, a naive treatment would give
938: a source term $\delta(t) \delta(m-M) \delta({\vec r}) / P(M)$ obtained by
939: simply dividing by $P(m)$, expressed at $m=M$ due to the Dirac
940: distribution $\delta(m-M)$.
941: However, this source term still depends on $m$ via the Dirac
942: distribution $\delta(m-M)$ and
943: is thus unsuitable as a source term of the equation
944: (\ref{thirdterter}) which is
945: independent of $m$. In order to circumvent this difficulty, one has to get
946: rid of the Dirac distribution $\delta(m-M)$. The corresponding
947: procedure has been
948: described in details in Ref.~\cite{HS1} and consists in applying the
949: integral operator $\int_{m_0}^{\infty} dm ~ {\hat \phi}(\beta,\vec r)$
950: to (\ref{third}), where ${\hat \phi}(\beta,\vec r)$ is the Laplace
951: transform with respect to the time variable of $\phi(t, \vec r)$.
952: In this way, the Dirac distribution $\delta(m-M)$ is regularized.
953: Identifying with the results of Ref.~\cite{HS1}, we obtain that
954: $N(t,\vec r)$ is
955: solution of (\ref{thirdterter}) with a source term
956: \be
957: S_M(t, \vec r)=\delta(r) \delta(t) \rho(M) /n ~,
958: \label{fngnfldq}
959: \ee
960: where $\rho(M)$ is defined in (\ref{formrho}) and $n$ is given by
961: (\ref{second}).
962: Thus, the complete Master equation for the
963: number $N(t,\vec r)$ of events at position $\vec r$ at time $t$ of any
964: possible magnitude is solution of
965: \be
966: N(t,\vec r) = S_M(t, \vec r) + \int d\vec r'
967: \int_0^t d\tau~ \phi(t-\tau,\vec r-\vec r') ~N(\tau, \vec r') ~, ~~~~t>0~,
968: \label{thirdter}
969: \ee
970: $N(t,\vec r)$ is the ``dressed'' or ``renormalized'' propagator, obtained
971: by summing the bare Omori propagator over all possible aftershock cascades.
972: $N(t,\vec r)$ can also be called the renormalized Omori law \cite{SS}.
973:
974: The essential assumption used to derive (\ref{third}) is that
975: the fluctuations of the earthquake magnitudes in a given sequence
976: can be considered to be decoupled from those of the seismic rate.
977: This approximation can be shown to be
978: valid for $\alpha \leq b/2$ \cite{HSG}, for which the random variable $\rho(m_i)$ has
979: a finite variance.
980: In this case, any coupling between the fluctuations of the earthquake energies
981: and the instantaneous seismic rate provides only sub-dominant corrections to the
982: equation (\ref{third}). For $\alpha > b/2$, the variance of $\rho(m_i)$ is
983: mathematically infinite or undefined as $\rho(m_i)$ is distributed according
984: to a power law with exponent $b/\alpha <2$.
985: In this case, the Master equation (\ref{third}) is not completely correct
986: as an additional term must be included to account for the effect of
987: the dependence between the fluctuations of earthquake magnitudes and
988: the instantaneous seismic rate. Our results
989: are presented below for $\alpha = 0.5$ which belongs to the first
990: regime $\alpha \leq b/2$. For $\alpha > b/2$, Ref.~\cite{HSG} has shown
991: that the renormalization of the bare propagator into the dressed propagator
992: is weaker than for $\alpha \leq b/2$, all the more so as $\alpha \to b$.
993: Preliminary numerical simulations for $\alpha > b/2$ shows that our results
994: presented below hold qualitatively but with a reduction of the observed
995: spatial diffusion exponent compared to the value predicted from the Master
996: equation approach developed here. This regime $\alpha > b/2$ is probably
997: relevant to the real seismicity \cite{YaShi,Guo2,H02}, even if a precise
998: estimation of $\alpha$ is very difficult.
999:
1000:
1001: \subsection{A Master equation of the CTRW model}
1002:
1003: We now demonstrate that the
1004: self-consistent mean field equation (\ref{thirdter}) is identical to
1005: the Master equation of a
1006: continuous-time random walk (CTRW). Random walks underlie many physical
1007: processes and are often the basis of first-order description of
1008: natural processes.
1009: The CTRW model, which is a generalization of the naive model of a random walker
1010: which jumps by $\pm 1$ spatial step on a discrete lattice at each
1011: time step, was introduced
1012: by \cite{Montroll1} and investigated by many other workers
1013: \cite{Montroll12,Scher,Kenkre,Shlesinger1,Weiss1}.
1014: The CTRW considers a continuous
1015: distribution of spatial steps as well as time steps (which can be seen
1016: either as waiting times between steps or as durations of the steps).
1017: The CTRW model is thus based on the idea that the length of a given
1018: jump, as well as the waiting time $\tau_i = t_i - t_{i-1}$
1019: elapsing between two successive jumps are drawn from a joint
1020: probability density function (pdf) $\phi(\vec r, t)$, which is
1021: usually referred to as the jump pdf. From a mathematical point of view,
1022: a CTRW is a process subordinated to
1023: random walks under the operational time defined by the process $\{t_i\}$.
1024:
1025: From $\phi(\vec r, t)$, the jump length pdf
1026: $\Phi(\vec r) = \int_0^{+\infty} dt~\phi(\vec r, t)$ and the waiting time
1027: pdf $\Psi(t) = \int d\vec r ~\phi(\vec r, t)$
1028: can be deduced. Thus, $\Phi(\vec r) d\vec r$ produces the probability
1029: for a jump length in the interval $(\vec r, \vec r+d\vec r)$
1030: and $\Psi(t) dt$ the probability for a waiting time in the interval
1031: $(t, t+dt)$.
1032: When the jump length and
1033: waiting time are independent random variables, this corresponds
1034: to the decoupled form $\phi(\vec r, t) = \Psi(t) ~ \Phi(\vec r)$.
1035: If both are coupled, a jump of
1036: a certain length involves a time cost or, vice versa in a given time
1037: span the walker can only
1038: travel a maximum distance. With these definitions, a CTRW process can
1039: be described
1040: through a Master equation (see \cite{Weiss1,Hughes,Meltzner2}
1041: for a review and references therein) which turns out to be given by an
1042: equation which is identical to (\ref{thirdter}).
1043:
1044: This connection between the ETAS model of earthquakes and a model of
1045: random walks provides an important advance for the understanding
1046: of spatio-temporal earthquake processes, as it allows one to borrow
1047: for the deep knowledge accumulated in past decades on random walks.
1048: In the same spirit,
1049: polymer physics acquired its status as a fundamental physical problem
1050: from its previous status of an applied field of research in chemistry
1051: when Flory, Edwards, de Gennes, des Cloizeaux and others showed how
1052: to formulate problems
1053: in polymer physics in the language of random walks and how to extract
1054: novel results. In the sequel of this article, we use this analogy
1055: to provide a wealth of new predictions as well as new questions for
1056: earthquake aftershocks.
1057:
1058: In the context of the CTRW, we have the following correspondence.
1059: \begin{itemize}
1060: \item $N(t,\vec r)$ is the pdf for the random walker to just
1061: arrive at position $\vec r$ at time $t$.
1062:
1063: \item The source term $S_M(t, \vec r)$ given by (\ref{fngnfldq})
1064: denotes the initial condition of the random walk, here chosen to be
1065: at the origin of space at time $t=0$. The constant $\rho(M) /n$ adds
1066: the possibility via the parameter $M$ to have more than one initial
1067: walker at the origin.
1068:
1069: \item In the CTRW context, the Master
1070: equation (\ref{thirdter}) states that the pdf $N(t,\vec r)$ of just
1071: having arrived at position $\vec r$ at time $t$ comes from all possible paths
1072: in number $N(\tau,\vec r')$
1073: having crossed a position $\vec r'$ at an earlier time $\tau$, weighted
1074: by a transfer or propagator function $\phi(t-\tau,\vec r-\vec r')$
1075: describing all the possible steps of the random walker from $(\tau,\vec r')$ to
1076: $(t,\vec r)$.
1077: \end{itemize}
1078:
1079: It is important to stress that $N(t,\vec r)$ defined above
1080: is different from the standard quantity $W(t,\vec r)$ usually studied
1081: in random walk problems,
1082: defined as the probability to find the random walk at position $\vec
1083: r$ at time $t$.
1084: The relationship between $N(t,\vec r)$ and $W(t,\vec r)$ is
1085: \be
1086: W(t,\vec r) = \int_0^t dt'~\left[ 1 - \int_0^{t-t'} dt''~
1087: \Psi(t'')\right]~N(t',\vec r)~.
1088: \label{bheiw}
1089: \ee
1090: The term $1 - \int_0^{t-t'} dt''~ \Psi(t'')$ in bracket is the probability for
1091: the walker not to jump in the time interval $[t',t]$ and the integral
1092: in the right-hand-side of (\ref{bheiw}) means that the probability
1093: $W(t,\vec r)$ for the random
1094: walker to be at position $\vec r$ at time $t$ is the sum over all possible
1095: scenarios in which the walker just arrives at $\vec r$ at an earlier
1096: time $t'$ and
1097: then does not jump until time $t$. In the context of earthquake aftershocks,
1098: $W(t,\vec r)$ is the probability that
1099: an event at $\vec r$ has occurred at a time $t' \leq t$
1100: and that the whole system has remained quiescent from $t'$ to $t$.
1101:
1102: In the Fourier-Laplace domain (see below), expression (\ref{bheiw}) reads
1103: \be
1104: {\hat W}(\beta,\vec k) = {1 - {\hat \Psi}(\beta) \over \beta}~{\hat
1105: N}(\beta,\vec k)~.
1106: \label{mhgjhjhd}
1107: \ee
1108:
1109:
1110:
1111: In general, the CTRW models transport phenomena in any
1112: heterogeneous media. It has for instance been used successfully for
1113: describing the behavior
1114: of chemical species as they migrate through porous media
1115: \cite{Margolin,Berko1}.
1116: In insight, it is rather natural that it can be applied to the
1117: ``transport of stress''
1118: through the heterogeneous crust and thus to the description of the anomalous
1119: diffusion of seismic activity.
1120:
1121: Table \ref{table1} synthesizes
1122: the correspondence between the ETAS and CTRW models and
1123: then draws its consequences.
1124:
1125:
1126:
1127: \subsection{Experimental verifications of the cross-over between the
1128: two power law Omori decays in photoconductivity in amorphous semi-conductors
1129: and in fractal stream chemistry
1130: using the correspondence between the ETAS and CTRW model}
1131:
1132: The crossover from an Omori law
1133: $1/t^{1-\theta}$ for $t<t^*$ to
1134: $1/t^{1+\theta}$ for $t>t^*$ found in
1135: \cite{SS,HS1} with $t^*$ given by (\ref{tstar})
1136: has actually a counterpart in the CTRW. This
1137: behavior was first studied by Scher and Montroll \cite{Scher} in a CTRW with
1138: absorbing boundary condition to model photoconductivity in amorphous
1139: semi-conductors
1140: As$_2$Se$_3$ and an organic compound TNF-PVK finding $\theta \approx 0.5$ and
1141: $\theta = 0.8$ respectively. In a semiconductor experiment, electric holes
1142: are injected near a positive electrode and then transported to a negative
1143: electrode where they are absorbed. The transient current follows exactly
1144: the transition $1/t^{1-\theta}$ for $t<t^*$ to
1145: $1/t^{1+\theta}$ for $t>t^*$ found for Omori law for earthquake aftershocks
1146: in the ETAS model. In the semiconductor context, the finiteness of $t^*$
1147: results from the existence of a force applied to the holes while in the ETAS
1148: model it results from a finite distance $1-n$ to the critical point $n=1$
1149: in the subcritical regime. When the force goes to zero or $n \to 1$,
1150: $t^* \to +\infty$.
1151:
1152: A similar transition has been recently proposed to model
1153: long-term time series measurements of chloride, a natural passive tracer,
1154: in rainfall and runoff in catchments \cite{Schermarklaber}. The
1155: quantity analogous to
1156: the dressed Omori propagator is the effective travel time distribution $h(t)$
1157: which governs the global lag time between injection of the tracer
1158: through rainfall
1159: and outflow to the stream. $h(t)$ has been shown to have a power-law form
1160: $h(t) \sim 1/t^{1-m}$ with $m$ between -0.3 and 0.2 for different time series
1161: \cite{Kirchner}. This variability may be due to the transition
1162: between an exponent
1163: $1-\theta$ at short times to $1+\theta$ at long times \cite{Schermarklaber},
1164: where $\theta$ is the exponent of the bare distribution of individual
1165: transition times.
1166:
1167:
1168: \subsection{General and formal solution of the spatial ETAS model}
1169:
1170: Let us solve (\ref{thirdter}) for the
1171: number $N(t,\vec r)$ of events at position $\vec r$ at time $t$ of any
1172: possible magnitude. Recall that $N(t,\vec r)$ can also be interpreted
1173: as the dressed Omori propagator. Extending \cite{HS1}
1174: to the spatial domain and also in analogy with the standard approach to
1175: solve the CTRW, the Laplace-in-time Fourier-in-space transform ${\hat
1176: N}(\beta,\vec k)$ of $N(t,\vec r)$ is given by
1177: \be
1178: {\hat N}(\beta,\vec k) = {{\hat S}_M(\beta,\vec k) \over
1179: 1 - n {\hat \Psi}(\beta) {\hat \Phi}(\vec k)}~,
1180: \label{ngnslqw}
1181: \ee
1182: where ${\hat S}_M(\beta,\vec k)$ is the Laplace Fourier transform of the source
1183: $S_M(t, \vec r)$ given by (\ref{fngnfldq})
1184: and ${\hat \Psi}(\beta)$ (respectively ${\hat \Phi}(\vec k)$) is the Laplace
1185: (respectively Fourier) transforms of $\Psi(t)$ (respectively $\Phi(\vec r)$).
1186: For a mainshock of magnitude $M$ occurring at time $t=0$ and position
1187: $\vec r=0$,
1188: the source term is thus ${\hat S}_M(\beta,\vec k)=\rho(M)/n$.
1189: The only difference between expression (\ref{ngnslqw}) and the Laplace-Fourier
1190: transform of the pdf of the CTRW of just having arrived at $\vec r$ at time $t$
1191: occurs when the branching ratio $n$ is different from $1$. In general,
1192: solutions of CTRW models are expressed for $n=1$ and for the variable
1193: $W(t, \vec r)$
1194: which is simply related to $N(t, \vec r)$ according to (\ref{bheiw}). Using
1195: (\ref{bheiw}) and (\ref{ngnslqw}) leads to
1196: \be
1197: {\hat W}(\beta,\vec k) = {1 - {\hat \Psi}(\beta) \over \beta}~~{{\hat
1198: S}_M(\beta,\vec k) \over
1199: 1 - n {\hat \Psi}(\beta) {\hat \Phi}(\vec k)}~,
1200: \label{ngnslaaqw}
1201: \ee
1202:
1203: In the following, we exploit (\ref{ngnslaaqw}) to
1204: obtain analytical solutions of the spatial ETAS model in
1205: different regimes,
1206: that provide specific predictions on the conditions necessary for observing
1207: aftershock diffusion. In addition, we provide specific predictions
1208: on the exponent $H$ of the
1209: diffusion law $R \sim t^{H}$ that are tested by numerical simulations.
1210:
1211:
1212: \section{Critical regime $n=1$}
1213:
1214: \subsection{Classification of the different regimes}
1215:
1216: Numerous works on the CTRW have investigated many possible forms for
1217: $\Psi(t)$ and $\Phi(\vec r)$ and have provided the asymptotic long time
1218: and large scale dependence of $W(t, \vec r)$ (see
1219: \cite{Weiss1,Hughes,Meltzner2,Berko1} and references therein).
1220: Here, we restrict our discussion to the cases where both $\Psi(t)$ and
1221: $\Phi(\vec r)$ have power law tails as given by (\ref{psidef}) and
1222: (\ref{phidef}).
1223: The long-time and large scale behavior of the ETAS and CTRW are controlled by
1224: the behavior of the Laplace-Fourier transforms for small $\beta$ and
1225: small $|\vec k|$.
1226:
1227: Two cases must be distinguished depending on the exponent $\mu$ controlling
1228: the weight of the tail of $\Phi(\vec r)$.
1229: \begin{itemize}
1230: \item For $\mu > 2$, the variance $\langle (\vec r)^2 \rangle =
1231: \sigma^2$ of the jump
1232: size distribution exists. To leading order in $k = |\vec k|$, ${\hat
1233: \Phi}(\vec k)$
1234: can be expanded as
1235: \ba
1236: {\hat \Phi}(\vec k) = 1- \sigma^2 k^2 + {\cal O}(k^o)~,~~~{\rm with}~~o>2~.
1237: \label{ngnglw}
1238: \ea
1239:
1240: \item For $\mu\leq 2$, the variance $\langle (\vec r)^2 \rangle$
1241: is infinite. This regime of ``long jumps'' leads to so-called L\'evy flights.
1242: In this case, to leading order in $k = |\vec k|$, ${\hat \Phi}(\vec k)$
1243: can be expanded as
1244: \be
1245: {\hat \Phi}(\vec k) = 1- \sigma^{\mu} k^{\mu} + {\cal O}(k^o)~,~~~{\rm where}~~
1246: 0 < \mu \leq 2, ~~~{\rm with}~~o>\mu~,
1247: \label{ngngfdlw}
1248: \ee
1249: where $\sigma$ is a characteristic distance defined by
1250: \be
1251: \sigma= \left\{ \begin{array}{lll}
1252: d~[\Gamma(1-\mu)]^{1/\mu} , & & 0<\mu < 1~, \\
1253: {d~\pi \over \mu ~ \Gamma(\mu-1)~\sin(\pi \mu /2) } , & & 1<\mu<2~.
1254: %\\
1255: %{2 d \over \sqrt{(\mu-1)(\mu-2)}} , & & \mu > 2~.
1256: \end{array}
1257: \right.
1258: \label{fklj}
1259: \ee
1260: \end{itemize}
1261:
1262: For a distribution $\Psi(t)$ of waiting times of the form
1263: of a local Omori law (\ref{psidef}) with exponent $\theta < 1$,
1264: ${\hat \Psi}(\beta)$ can be expanded for small $\beta$ as
1265: \be
1266: {\hat \Psi}(\beta) = 1 - (\beta c')^{\theta} + {\cal
1267: O}(\beta^\omega)~,~~~{\rm with}~
1268: \omega \geq 1~.
1269: \label{hgngwlff}
1270: \ee
1271: where $c'$ is proportional to $c$ up to a numerical constant
1272: $c'=c \left(\Gamma(1-\theta)\right)^{1/\theta}$ in the case $\theta<1$.
1273:
1274: Putting the leading terms
1275: of the expansions of ${\hat \Phi}(\vec k)$ for small $|\vec k|$ and
1276: of ${\hat \Psi}(\beta)$
1277: for small $\beta$ in (\ref{ngnslqw}) gives
1278: \be
1279: {\hat N}(\beta,\vec k) = {{\hat S}_M(\beta,\vec k) \over 1-n +
1280: n(\beta c')^{\theta} + n\sigma^{\mu} k^{\mu}}~.
1281: \label{ngnslqwassa}
1282: \ee
1283: The corresponding ${\hat W}(\beta,\vec k)$ is obtained from
1284: (\ref{ngnslaaqw}) by
1285: \be
1286: {\hat W}(\beta,\vec k) = {\hat S}_M(\beta,\vec k)~{(\beta)^{\theta
1287: -1} c'^{\theta} \over
1288: 1-n +n (\beta c')^{\theta} + n\sigma^{\mu} k^{\mu}}~.
1289: \label{ngnslqssaawaa}
1290: \ee
1291:
1292:
1293: The critical regime $n=1$ gets rid of the constant term $1-n$ in the
1294: denominator of (\ref{ngnslqwassa}) and (\ref{ngnslqssaawaa}). This case
1295: is analyzed in details below.
1296:
1297: The regime $n\neq 1$ introduces a characteristic time $t^*$ given by
1298: (\ref{tstar}).
1299: In the sub-critical regime, equation (\ref{ngnslqwassa}) can be rewritten as
1300: \be
1301: {\hat N}(\beta,\vec k) = {{\hat S}_M(\beta,\vec k) \over (1-n)}~
1302: {1 \over 1 + (\beta t^*)^{\theta} + (k r^* )^{\mu}}~.
1303: \label{ngnslqwassa2}
1304: \ee
1305: where $r^*$ is defined by
1306: \be
1307: r^*=\sigma \left({n \over 1-n}\right)^{1/\mu}~.
1308: \label{gnjgrkd}
1309: \ee
1310: For $t<t^*$ and $r<r^*$, the dressed propagator is given by the same
1311: expression as
1312: for the critical case and all our results below hold.
1313: For large times $t>t^*$ and large distances $r>r^*$,
1314: we can factorize (\ref{ngnslqwassa2}) as a
1315: product of a function of time and a function of space
1316: \be
1317: {\hat N}(\beta,\vec k) \simeq {{\hat S}_M(\beta,\vec k) \over (1-n)}
1318: {1 \over (1+(\beta t^*)^{\theta})}{1 \over (1 + (k r^*)^{\mu})}~.
1319: \label{ngnslqwassa3}
1320: \ee
1321: Thus, there is no diffusion in the sub-critical regime for $t>t^*$ and $r>r^*$.
1322: We shall not analyze further this
1323: trivial regime $n<1$ and $t>t^*$ and will only analyze the case $t<t^*$.
1324: If there is the need, the cross-over can be calculated
1325: explicitly using (\ref{ngnslqwassa}).
1326:
1327: In order to get the leading behavior of $N(t, \vec r)$ from that of
1328: $W(t, \vec r)$,
1329: we see from (\ref{ngnslqw}) and (\ref{ngnslaaqw}) that
1330: ${\hat N}(\beta,\vec k) = {\beta \over 1 - {\hat \Psi}(\beta)}~{\hat
1331: W}(\beta,\vec k)
1332: \approx \beta^{1-\theta} c'^{-\theta} ~{\hat W}(\beta,\vec k)$. The
1333: inverse Laplace transform
1334: of $1/\beta^{\theta}$ is $1/[\Gamma(\theta)~t^{1-\theta}]$. Using the
1335: fact that the Laplace transform of $df/dt$ is $\beta$ times the
1336: Laplace transform
1337: of $f(t)$ minus $f(0)$, we get $N(t, \vec r)$ as the
1338: derivative of a convolution
1339: \be
1340: N(t, \vec r) = {c'^{-\theta} \over \Gamma(\theta)}
1341: {d \over dt}~\int_0^t dt'~ {W(t', \vec r) \over (t-t')^{1-\theta}}
1342: = c'^{-\theta}~ _0D_t^{1-\theta}~ W(t, \vec r)~.
1343: \label{nbjn}
1344: \ee
1345: In (\ref{nbjn}), we have dropped the Dirac function coming from the inverse
1346: Laplace transform of the constant term $f(0)$, which provides
1347: a contribution only at the origin of time $t=0$. Note that the
1348: operator ${1 \over \Gamma(\theta)}
1349: {d \over dt}~\int_0^t dt'~ {W(t', \vec r) \over (t-t')^{1-\theta}}$
1350: is nothing but the so-called fractional Riemann-Liouville derivative operator
1351: of order $1-\theta$ applied to the function $W(t, \vec r)$ of time $t$ and is
1352: usually denoted
1353: $_0D^{1-\theta}_t W(t, \vec r)$.
1354:
1355:
1356: \subsection{The standard diffusion case $\theta >1$ and $\mu>2$ \label{mgmls}}
1357:
1358: The standard diffusion process is recovered for $\theta \geq 1$
1359: (for which the average waiting time is finite)
1360: and for $\mu \geq 2$ (for which the variance of the jump length is finite).
1361: In this case,
1362: ${\hat N}(\beta,\vec k) = {{\hat S}_M(\beta,\vec k) \over
1363: \beta c' + \sigma^{2} k^{2}}$. For an impulsive source
1364: leading to ${\hat S}_M(\beta,\vec k) =$ constant, this is the
1365: Laplace-Fourier transform of the standard diffusion propagator
1366: \be
1367: N(t, \vec r) \propto {1 \over (D t)^{d/2}}~
1368: \exp [ - (\vec r)^2/Dt]~,~~~~{\rm where}~~D = \sigma^{2}/c'~,
1369: \label{gngk}
1370: \ee
1371: where $d$ is here the space dimension.
1372: This solution is valid for $|\vec r|/\sqrt{Dt}$ not too large. For
1373: larger values,
1374: large deviations lead to corrections with the power law tail of the
1375: input jump distribution $\Phi(\vec r) \sim 1/|\vec r|^{1+\mu}$ defined in
1376: (\ref{phidef}), along the lines presented for instance in
1377: \cite{Sorbook} (section 3.5). This regime is not relevant to the
1378: aftershock problem for which usually $0 < \theta <1$.
1379:
1380:
1381: \subsection{Long waiting times ($\theta < 1$) and finite variance
1382: of the jump sizes ($\mu>2$)}
1383:
1384: Putting the leading terms of the expansions of ${\hat \Phi}(\vec k)$
1385: (\ref{ngnglw}) and of ${\hat \Psi}(\beta)$ (\ref{hgngwlff})
1386: in (\ref{ngnslqw}) gives
1387: \be
1388: {\hat N}(\beta,\vec k) = {1 \over (\beta c')^\theta + (\sigma k)^2}
1389: \label{ngbnmvd}
1390: \ee
1391:
1392: The expression (\ref{ngbnmvd}) can be inverted with respect to the
1393: Fourier transform, and then
1394: inverted with respect to the Laplace transform using Fox functions
1395: \cite{Meltzner2,barkai3}.
1396: The solution for $W(t, \vec r)$ in one dimension is given for instance
1397: in \cite{Meltzner2} in terms of an infinite sum
1398: % eq 46 de [{\it Metzler and Klafter}, 2000]
1399: \be
1400: W(t,\vec r) ={ 1 \over 2 D}~{1 \over t^{\theta \over 2}}~
1401: \sum_{k=0}^{\infty} {(-1)^k ~ z^{-k} \over k!~\Gamma(1-\theta (k+1)/2)}
1402: \label{giouyt}
1403: \ee
1404: where
1405: \be
1406: z= { D~t^{\theta/2} \over |\vec r|}
1407: \label{zdef}
1408: \ee
1409: and $D=\sigma /c'^{\theta/2}$. % metzler and klafter eq 41
1410:
1411: Expression (\ref{giouyt}) and many others below involve the Gamma function of
1412: negative arguments. We recall that the Gamma function $\Gamma(u)$ can be
1413: analytically continued to the whole complex plane, except for the simple poles
1414: $u=0, -1, -2, -3, ...$ Thus, $\Gamma(u)$
1415: is defined everywhere but at these poles.
1416: In order to get the expression of the Gamma
1417: function for negative arguments, one can use two formulae:
1418: $\Gamma(1-u) \times \Gamma(u) = \pi /\sin(\pi u)$ and
1419: $\Gamma(1+u) = u \Gamma(u)$.
1420: Both these formulae are valid for all points with the possible exception of
1421: the arguments at poles $0, -1, -2,...$ For instance,
1422: $\Gamma(-\theta) = \Gamma(1-\theta) / (-\theta) =
1423: -[\pi /\theta \sin(\pi \theta)]/ \Gamma(\theta)$, for $0 < \theta < 1$.
1424:
1425: Expression (\ref{giouyt}) can be rewritten as a Fox-function \cite{Mathai}
1426: \be
1427: W(t,z) = {1 \over 2D}~ {1 \over t^{\theta \over 2}}~ H_{1,1}^{1,0}
1428: \left[{1 \over z} \left |
1429: {\begin{array}{l} (1-\theta/2,\theta/2) \\ (0,1) \end{array}}
1430: \right .
1431: \right ]
1432: \label{fox4346}
1433: \ee
1434: whose asymptotic dependence for large $z$,
1435: obtained from a standard theorem of the Fox function (equation (1.6.3) of
1436: \cite{Mathai}),
1437: \be
1438: W(t, z) \sim {1 \over D~t^{\theta \over 2}}~
1439: {1 \over z^{1-\theta \over 2-\theta}}~
1440: \exp \left(-\left(1-{\theta \over 2}\right)\left({\theta \over
1441: 2}\right)^{\theta \over 2- \theta}
1442: z^{2 \over 2-\theta} \right )
1443: \label{exp56}
1444: \ee
1445: is in agreement with the result of Roman and Alemany \cite{Roman} and
1446: Barkai et al. \cite{barkai3}
1447: for a space dimension $d_f=1$,
1448: including the dependence in the power law prefactor to the exponential.
1449: The exponential dependence $W(t, r) \sim
1450: \exp \left(- {\rm const}~ (r/D t^{\theta/2})^{2 \over 2-\theta} \right )$
1451: in (\ref{exp56}) holds in arbitrary dimensions $d_f$, the only
1452: modification occurring
1453: in the prefactor whose power of $z$ change with the space
1454: dimension $d_f$ as \cite{Roman,barkai3}
1455: \be
1456: W_{d_f}(t, z) \sim {1 \over D~ t^{{\theta \over 2}}}~
1457: {1 \over z^{d_f (1-\theta) \over 2-\theta}}~
1458: \exp \left(-\left(1-{\theta \over 2}\right)\left({\theta \over
1459: 2}\right)^{\theta \over 2- \theta}
1460: z^{2 \over 2-\theta} \right )~.
1461: \label{exp5a6}
1462: \ee
1463:
1464: The expression of $N(t,\vec r)$ can be obtained from $W(t,\vec r)$
1465: using the fractional
1466: Riemann-Liouville derivation (\ref{nbjn}) of order $1-\theta$.
1467: Inserting expression (\ref{giouyt}) in (\ref{nbjn})
1468: and using the expression of
1469: the fractional Riemann-Liouville derivative operator
1470: $_0D^{\alpha}_t$ applied to an arbitrary power $t^{\mu}$, i.e.,
1471: $_0D^{\alpha}_t t^{\mu} = {\Gamma (1+\mu) \over
1472: \Gamma(1+\mu-\alpha)}~t^{\mu-\alpha}$, we obtain
1473: % equation A.16 de [Metzler and Klafter, 2000]
1474: % see eq A.16 of metzler and klafter for the fractional derivative
1475: % of a power law
1476: \be
1477: N(t,\vec r) = {c'^{-\theta} \over 2 D t^{1-{\theta \over 2}}}~
1478: \sum_{k=0}^{\infty} {(-1)^k ~ z^k \over k!~ \Gamma((1-k)\theta /2)}~.
1479: \label{qdsi}
1480: \ee
1481: Expression (\ref{qdsi}) can be used to evaluate $N(t,\vec r)$ for small $z$,
1482: but the numerical evaluation of (\ref{qdsi}) is impossible for large $z$.
1483: In order to obtain the asymptotic behavior of $N(t,\vec r)$,
1484: expression (\ref{qdsi}) can be rewritten as a Fox-function \cite{Mathai}
1485: \be
1486: N(t,\vec r) = {c'^{-\theta} \over 2D t^{1-{\theta \over 2}}}~ H_{1,1}^{1,0}
1487: \left[ {1 \over z } \left |
1488: {\begin{array}{l} (\theta/2,\theta/2) \\ (0,1) \end{array}}
1489: \right .
1490: \right ] ~.
1491: \label{fox3}
1492: \ee
1493: Employing again the standard theorem of the Fox function (equation (1.6.3) of
1494: \cite{Mathai}),
1495: the asymptotic behavior of $N(t, r)$ for large distances $r$
1496: such that $r>D t^{\theta/2}$ is given by
1497: \be
1498: N(t, r) \sim {c'^{-\theta} \over D t^{1-{\theta \over 2}}}~
1499: \left({|\vec r| \over D t^{\theta/2}} \right)^{1-\theta \over 2-\theta}~
1500: \exp \left(-\left(1-{\theta \over 2}\right)\left({\theta \over
1501: 2}\right)^{\theta \over 2- \theta}
1502: \left({|\vec r| \over D t^{\theta/2}} \right)^{2 \over 2-\theta} \right )~.
1503: \label{exp1}
1504: \ee
1505: The exponential dependence $N(t, r) \sim
1506: \exp \left(- {\rm const}~ (r/D t^{\theta/2})^{2 \over 2-\theta} \right )$
1507: in (\ref{exp1}) holds in arbitrary dimensions.
1508:
1509: This expression becomes incorrect for very large distances because it would
1510: predict an exponential or slightly super-exponential decay with $r$.
1511: This cannot be true as the global law cannot decay faster than the
1512: local law (\ref{phidef}). The reason for (\ref{exp1}) to become incorrect at
1513: large distances is that the expansion of
1514: ${\hat N}(\beta,\vec k)$ for small $|\vec k|$ (large distances) given by
1515: (\ref{ngbnmvd}) has been truncated at the order $k^{2}$. There is however
1516: a subdominant term $\propto k^{\mu}$ that describes the power law tail
1517: of the local law (\ref{phidef}) and also of the global law asymptotically.
1518: A similar situation occurs in the application of the
1519: central limit theorem for sums of $N$ random variables with power law
1520: distributions with exponents $\mu>2$ \cite{Sorbook}: the distribution
1521: of the sum $S$
1522: is a Gaussian in its bulk for $|S| < \sqrt{N \ln N}$ and crosses over
1523: to a power law
1524: with tail exponent $\mu$ for larger $S$. In a similar way, the cross-over
1525: of $N(t, r)$ to the
1526: asymptotic local power law (\ref{phidef}) can be recovered by an analysis
1527: including the subleading correction $\propto k^{\mu}$ to the expansion
1528: (\ref{ngbnmvd}).
1529:
1530: Expression (\ref{qdsi}) shows that the global rate of seismicity cannot be
1531: factorized as a product of a distribution of times and a distribution
1532: of distances.
1533: This space-time coupling implies that the seismic activity diffuses with time,
1534: and that the decay
1535: of the rate of aftershocks depends on the distance from the first mainshock.
1536: This coupling of space and time stems from the cascade of
1537: aftershocks, from the primary aftershocks to the secondary aftershocks to the
1538: tertiary aftershocks and so on.
1539:
1540: Figure \ref{nrtmus2} presents the decay of the seismic activity
1541: $N(r,t)$ obtained
1542: using expression (\ref{qdsi}) for small $z$ and expression
1543: (\ref{exp1}) for large $z$,
1544: as a function of the time from the mainshock and as a function of the
1545: distances $r$. Close to the mainshock epicenter, expression (\ref{qdsi})
1546: predicts that the global seismicity rate decays with time as the
1547: renormalized Omori law
1548: \be
1549: N(t,0) \sim {1 \over t^{1-\theta/2}}~.
1550: \label{oyu}
1551: \ee
1552: The same decay is found at any fixed point $\vec r$ for times $t >
1553: (|\vec r|/D)^{2/\theta}$. At all times, the same decay $1/t^{1-\theta/2}$
1554: is also obtained by measuring the
1555: aftershock seismicity in a local box at a distance from the main shock origin
1556: increasing with time as $r \sim t^{\theta \over 2}$ (this is nothing
1557: but putting $z=$ constant in (\ref{qdsi})).
1558: At large distances $r > D t^{\theta/2}$, the global decay law is
1559: different from a
1560: power-law decay. Figure \ref{nrtmus2} shows that the rate of
1561: aftershocks presents
1562: a truncation at early times, which increases as the distance $r$ increases.
1563: At large times, the rate of
1564: aftershocks recovers the $1/t^{1-\theta/2}$ power-law decay (\ref{oyu}).
1565: We stress that a fit of the global law $N(r,t)$ over the whole
1566: time interval by an Omori law would yield an apparent
1567: exponent $p< 1-\theta/2$ that decreases with $r$.
1568:
1569: Integrating (\ref{qdsi}) over the whole one-dimensional space, we recover the
1570: global Omori law
1571: \be
1572: N(t)=\int dr N(t,r) \sim {1 \over t^{1-\theta}}
1573: \ee
1574: found in \cite{SS,HS1}. Thus, we have found an additional source of variability
1575: of the exponent $p$ of the Omori law: if measured over the whole catalog,
1576: we should measure $p=1-\theta$ in the critical regime $n=1$ while
1577: $p=1-\theta/2$ is
1578: slightly larger when measured in certain time- and space-windows,
1579: as described above. Thus, in this regime,
1580: pruning of catalogs may lead to continuous change
1581: from the value $1-\theta$ to $1-\theta/2$. In addition, as we have mentioned,
1582: the cross-over in time may lead to still smaller apparent exponents, thus
1583: enhancing the impression of variability of the exponent $p$. In reality, this
1584: range of $p$-values are seen to result from the complex spatio-temporal
1585: organization of the aftershock seismicity of the ETAS model. These results
1586: should lead us to be cautious when analyzing real catalogs with respect
1587: to the conditions and regimes under which the analysis is performed.
1588:
1589: There is another observable that
1590: characterizes how an aftershock sequence invades space as a function of time.
1591: Expression (\ref{qdsi}) indeed predicts a sub-diffusion process quantified by
1592: \be
1593: \langle |\vec r|^2 \rangle \sim t^{2H}~, \label{hgnvaa}
1594: \ee
1595: with $H = \theta/2$ since the natural variable is
1596: $z$ given by (\ref{zdef}).
1597: Indeed, expression (\ref{qdsi})
1598: tells us that, up to a global rescaling function of time, the rate of
1599: aftershocks is identical for a fixed value of $z$. Thus, any
1600: aftershock structure
1601: diffuses according to (\ref{hgnvaa}).
1602:
1603: This prediction is checked in Figure \ref{poisq} by numerical simulations.
1604: 1000 synthetic catalogs have been generated with $\mu=3$, $\theta=0.2$
1605: and $n=1$. The average distance between the first mainshock
1606: and its aftershocks as a function of the time from the mainshock has
1607: been averaged over these 1000 simulations.
1608: The theoretical diffusion exponent is $H=\theta/2=0.1$, in
1609: good agreement with the
1610: asymptotic behavior observed in the numerical simulation.
1611: In practice, in order to minimize the effect of fluctuations and optimize
1612: the speed of convergence, we estimate numerically $\exp [\langle \ln
1613: |\vec r| \rangle]$ which is also expected to scale as
1614: $\exp [\langle \ln |\vec r| \rangle] \sim t^{\theta/2}$
1615: due to the simple scaling form of (\ref{fox3}).
1616:
1617: This problem has also been solved exactly in \cite{barkai1} in the context
1618: of the so-called fractional Fokker-Planck equation, which amounts to
1619: replace the distribution $\Phi(\vec r)$ of jumps (\ref{phidef}) by a
1620: Gaussian function.
1621: This fractional Fokker-Planck equation allows one to introduce the
1622: possibility of bias
1623: or drift in the CTRW and therefore in the aftershock sequence.
1624:
1625:
1626: \subsection{Exponential waiting time distribution and long jump size
1627: L\'evy distribution ($\mu<2$)}
1628:
1629: This case with an exponential distribution
1630: \be
1631: \Psi(t) = \lambda ~e^{-\lambda t}
1632: \label{ngllsa}
1633: \ee
1634: of waiting times with a L\'evy distribution
1635: $\Phi(\vec r)= L_{\mu}(|\vec r|)$ of jump sizes with tail exponent $\mu<2$
1636: has been investigated by Budde et al. \cite{Budde}.
1637: One finds
1638: \be
1639: \langle |\vec r|^2 \rangle^{1/2} \sim t^{1/\mu}~, \label{hgnvaaaa}
1640: \ee
1641: corresponding to a superdiffusion regime with Hurst exponent $H=1/\mu >1/2$.
1642: The full distribution function $W(t, \vec r)$ corresponding to the
1643: critical regime $n=1$ is known for $\lambda t >> 1$:
1644: \be
1645: W(t, \vec r) \propto {1 \over (\lambda t)^{1/\mu}}~ L_{\mu}\left({|\vec r|
1646: \over (\lambda t)^{1/\mu}}\right)~.
1647: \label{mgmkrl}
1648: \ee
1649: The corresponding $N(t, \vec r)$ is obtained from (\ref{mhgjhjhd}).
1650: The Laplace transform of the exponential distribution (\ref{ngllsa})
1651: is ${\hat \Psi}(\beta)= \lambda/(\beta + \lambda)$. We thus get
1652: \be
1653: {\hat N}(\beta,\vec k) =
1654: \left( \beta + \lambda \right) ~{\hat W}(\beta,\vec k)~,
1655: \ee
1656: and thus
1657: \be
1658: N(t, \vec r) = {\partial W(t, \vec r) \over \partial t}~+~\lambda
1659: ~W(t, \vec r)~.
1660: \label{mgkjhgk}
1661: \ee
1662: Expression (\ref{mgkjhgk}) together with (\ref{mgmkrl}) predicts
1663: a diffusion law $r \sim t^H$ with $H=1/\mu$ which is in good agreement
1664: with our simulations.
1665: At large times $|{\vec r}| \ll (\lambda t)^{1/\mu}$,
1666: $N(t, \vec r) \approx \lambda ~W(t, \vec r) \sim 1/t^{1/\mu}$, given
1667: an apparent local Omori exponent $\theta = 1 - 1/\mu$.
1668: This offers a new mechanism for generating Omori law for aftershocks
1669: from purely
1670: exponential local relaxation but with a heavy distribution of jump sizes.
1671: This power-law decay should be observed only at a fixed distance $r$
1672: or over a limited domain from the mainshock
1673: in the regime of large times.
1674:
1675: Integrating over the whole space, $\int d\vec r ~W(t, \vec r) =1$ which gives
1676: $N(t) = \delta(t) + \lambda$ equal to a constant seismic rate. This
1677: results from an initial mainshock at $t=0$ leading to the cascade of
1678: aftershocks
1679: adjusting delicately to this constant rate for the critical value $n=1$ of
1680: the branching parameter. In the sub-critical regime $n<1$, the Omori law
1681: integrated over space gives instead $N(t) \propto \exp [-(1-n)\lambda t]$,
1682: showing that the characteristic decay time $1/(1-n)\lambda$ of the dressed
1683: Omori propagator $N(t)$
1684: becomes much larger (much longer memory) that the decay time
1685: $1/\lambda$ of the bare
1686: Omori propagator.
1687:
1688: For $\mu>2$, we recover the
1689: standard diffusion corresponding to $\theta >1$ and $\mu>2$ discussed
1690: in section \ref{mgmls}.
1691:
1692:
1693: \subsection{Long waiting times ($\theta < 1$) and long jump sizes
1694: (L\'evy flight regime for $\mu \leq 2$)}
1695:
1696: Putting the leading terms of the expansions of ${\hat \Phi}(\vec k)$
1697: and of ${\hat \Psi}(\beta)$ in (\ref{ngnslqw}) gives
1698: \be
1699: {\hat N}(\beta,\vec k) = {\hat S}_M(\beta,\vec k)~{1 \over
1700: (\beta c')^{\theta} + (\sigma k)^{\mu}}~.
1701: \label{ngnslqwaa}
1702: \ee
1703: The corresponding ${\hat W}(\beta,\vec k)$ is given by
1704: \be
1705: {\hat W}(\beta,\vec k) = {\hat S}_M(\beta,\vec k)~{(\beta)^{\theta
1706: -1} c'^{\theta} \over
1707: (\beta c')^{\theta} + (\sigma k)^{\mu}}~.
1708: \label{ngnslqaawaa}
1709: \ee
1710:
1711: Equation (\ref{ngnslqaawaa}) has been studied extensively in the context
1712: of the CTRW model as a long wavelength $|\vec k| \to 0$ and long time
1713: $\beta \to 0$
1714: approximation to investigate the long time behavior of the CTRW.
1715: Kotulski \cite{Kot1} has developed a rigorous approach, based on limit
1716: theorems, to classify
1717: the asymptotic behaviors of different type of CTRWs and justifies the
1718: approximation
1719: (\ref{ngnslqaawaa}) for the long time behavior. Barkai \cite{barkai2}
1720: has studied
1721: the quality of the long wavelength $|\vec k| \to 0$ and long time $\beta \to 0$
1722: approximation (\ref{ngnslqaawaa}) by solving the exact CTRW problem for
1723: the case when the waiting time distribution $\Psi(t)$ is a one-sided
1724: stable L\'evy law of index $\theta$ with the same tail as (\ref{psidef})
1725: and the distribution $\Phi(\vec r)$ of jumps is a symmetric
1726: stable L\'evy of index $\mu$ with the same tail as (\ref{phidef}).
1727: Their Laplace and Fourier transforms,
1728: that appear in the denominator of (\ref{ngnslaaqw}), are
1729: respectively ${\hat \Psi}(\beta) = \exp [-\beta^{\theta}]$ and
1730: ${\hat \Phi}(\vec k) = \exp [ -|\vec k|^{\mu}/2 ]$. Note that the
1731: long wavelength $|\vec k| \to 0$ and long time $\beta \to 0$ approximation
1732: gives $1-\exp [-(c' \beta)^{\theta}] ~\exp [ -|\sigma \vec k|^{\mu}]=
1733: (c' \beta)^{\theta}+|\sigma \vec k|^{\mu}$, which recovers (\ref{ngnslqwaa}).
1734: By comparing the exact solution of (\ref{ngnslqw}) for $\Psi(t)$ and
1735: $\Phi(\vec r)$ of the above L\'evy form with that of
1736: the long wavelength $|\vec k| \to 0$ and long time $\beta \to 0$
1737: approximation (\ref{ngnslqaawaa}), Barkai \cite{barkai2} finds that certain
1738: solutions of (\ref{ngnslqaawaa}) diverge on the origin, a behavior not found
1739: for the corresponding solutions of (\ref{ngnslqw}). In addition, certain
1740: solutions of the full equation (\ref{ngnslqw}) converge only very
1741: slowly for $\mu <1$
1742: to the solutions of the long-time approximation (\ref{ngnslqaawaa}). These
1743: results validate our use of the asymptotic long time behavior with respect
1744: to the scaling laws but provide a note of caution if one needs more precise
1745: non-asymptotic information. In this case, such information can be obtained
1746: by a suitable analysis of the full equation (\ref{ngnslqw}).
1747:
1748: Using power counting,
1749: expression (\ref{ngnslqaawaa}) predicts a diffusion process
1750: (\ref{hgnvaa}) with exponent
1751: \be
1752: H = {\theta \over \mu}~.
1753: \label{mgjgkr}
1754: \ee
1755: This prediction is checked by numerical simulation of the ETAS model
1756: in the critical regime $n=1$, with $\theta=0.2$, $\mu=0.9$, shown in figure
1757: \ref{poimlksq}. The average distance between the first mainshock and its
1758: aftershocks as a function of the time from the mainshock indeed increases
1759: according to (\ref{hgnvaa}) with an exponent $H$ in very
1760: good agreement with the prediction $H=\theta/ \mu=0.22$. As
1761: the form of the denominator in (\ref{ngnslqaawaa}) is
1762: independent of the space dimension, the prediction
1763: (\ref{mgjgkr}) is valid in any space dimension.
1764:
1765: The natural variable for the expansions given below allowing to compute
1766: $N(t, \vec r)$ is
1767: \be
1768: z={ D ~t^{\theta/\mu} \over |\vec r|}~,
1769: \label{mgjgrk}
1770: \ee
1771: where $D=\sigma / c'^{\theta/\mu}$ and $c'=c
1772: \left(\Gamma(1-\theta)\right)^{1/\theta}$.
1773:
1774:
1775: \subsubsection{$z$-expansion of the solution}
1776:
1777: $W(t, \vec r)$ can be obtained as the following sum (equation (5.10)
1778: of \cite{Saichev})
1779: \be
1780: W(t, \vec r) = {1 \over \pi |\vec r|}~
1781: \sum_{m=0}^{+\infty} (-1)^m~ z^{m \mu}
1782: ~{\Gamma(m \mu +1) \over \Gamma(m \theta +1)} ~\cos \left[ {\pi \over
1783: 2} (m \mu +1)\right]~.
1784: \label{njhgrjb}
1785: \ee
1786:
1787: Applying (\ref{nbjn}) to (\ref{njhgrjb}) term by term in the sum, we get
1788: \be
1789: N(t, \vec r) = {c'^{-\theta} \over D \pi ~
1790: t^{1-\theta+\theta/\mu}}~
1791: \sum_{m=0}^{+\infty} (-1)^m~ z^{1+m \mu}
1792: ~{\Gamma(m \mu +1) \over \Gamma((m+1)\theta)} ~
1793: \cos \left[ {\pi \over 2} (m \mu +1)\right]~,
1794: \label{njhaagrssjb}
1795: \ee
1796: The asymptotics
1797: \be
1798: {\Gamma(m \mu+\mu +1)~\Gamma(m \theta +1)
1799: \over \Gamma(m \theta + \theta +1)~\Gamma(m \mu +1)} \sim
1800: {\Gamma(m \mu +\mu+1)~\Gamma((m+1)\theta) \over \Gamma((m+2)\theta)~
1801: \Gamma(m \mu +1)}
1802: \sim m^{\mu - \theta}
1803: \label{mgjgl}
1804: \ee
1805: show that the series (\ref{njhgrjb})
1806: and (\ref{njhaagrssjb}) exist only for $\mu < \theta$. It can be shown that
1807: these series exist for all $z$ in this case.
1808: This series converges very slowly for large $z$ but
1809: the Pad\'e summation method \cite{BeOr} can be used to improve the convergence
1810: of (\ref{njhaagrssjb}) in the case $\mu < \theta$, and can also be used
1811: to estimate (\ref{njhaagrssjb}) in the case $\mu > \theta$ for which the
1812: series diverges.
1813:
1814: The space integral $\int dr ~N(t, r)$ over the whole one-dimensional
1815: volume $V$,
1816: with $N(t, r)$ given by (\ref{njhaagrssjb}),
1817: recovers the global Omori law
1818: \be
1819: \int_V dr ~N(t, r) \sim {1 \over t^{1 - \theta}}~.
1820: \label{nhnkjbkf}
1821: \ee
1822: Note the non-trivial phenomenon in which the superposition of all
1823: aftershock activities transforms
1824: the local Omori law or ``bare propagator'' (\ref{psidef})
1825: $\Psi(t) \sim {1 \over t^{1+\theta}}$ into the global
1826: Omori law or ``dressed propagator'' ${1 \over t^{1 - \theta}}$.
1827: This effects was predicted in \cite{SS,HS1} in the version of the ETAS model
1828: without space dependence.
1829: These results are consistent with the claim of section
1830: \ref{consosk} according to which all results
1831: reported previously for the version
1832: of the ETAS model without space dependence hold also for the version of the
1833: space-dependent ETAS model studied here, when averaging over the whole space.
1834:
1835: The asymptotic behavior for $|\vec r| \gg D~ t^{\theta \over \mu}$
1836: (i.e., $z \ll 1$)
1837: and $\mu < \theta$ is obtained by keeping only the first non-zero
1838: term ($m=1$) in (\ref{njhaagrssjb})
1839: which is convergent for all $z$ in the case $\mu < \theta$
1840: \be
1841: N(t, \vec r) = { \sin \left({\pi \mu \over 2}\right) \over \sigma
1842: c'~\pi}~ {\Gamma(1+\mu)
1843: \over \Gamma(2\theta)}~\left({c' \over t}\right)^{1-2 \theta}~
1844: \left({\sigma \over |\vec r|}\right)^{1+\mu}~,~~~~{\rm for}~~
1845: |\vec r| \gg D~t^{\theta \over \mu}~.
1846: \label{gjjsls}
1847: \ee
1848: At fixed large $|\vec r|$ and for $t< |{\vec r}/D|^{\mu \over \theta}$,
1849: this predicts a local Omori law with exponent $p=1-2\theta$.
1850:
1851:
1852: \subsubsection{$1/z$-expansion of the solution}
1853:
1854: We use the theory of Fox functions \cite{Mathai} to obtain
1855: $N(t,\vec r)$ as an infinite series in $1/z$. For this, we first
1856: rewrite expression (\ref{njhaagrssjb}) as a Fox
1857: function \cite{Mathai}
1858: \be
1859: N(t,\vec r) = {c'^{-\theta} \over D ~\mu~ \pi~
1860: ~ t^{1-\theta+\theta/\mu}}~
1861: R\left( H_{2,2}^{1,2}
1862: \left[ z~e^{i \pi /2} \left |
1863: {\begin{array}{l} (1/\mu, 1/\mu),(1,1) \\ (1/\mu, 1/\mu),
1864: (\theta/\mu-\theta+1, \theta/\mu) \end{array}}
1865: \right .
1866: \right ] \right)~,
1867: \label{fox4}
1868: \ee
1869: where $R(z)$ indicates the real part of $z$.
1870:
1871: The $1/z$ expansion of $N(t, \vec r)$ can be obtained using
1872: the dual expansion of the Fox function (\ref{fox4}) (expression
1873: (3.7.2) of \cite{Mathai})
1874: $$
1875: N(t,\vec r) = {c^{-\theta} \over D ~ \pi~ \mu~
1876: ~ t^{1-\theta+\theta/\mu}}~
1877: \sum_{m=0}^{+\infty} (-1)^m ~\biggl[\mu~
1878: z^{1-\mu-m\mu}~{\Gamma(1-(m+1)\mu) ~\sin((m+1)\mu\pi/2)
1879: \over \Gamma(-m \theta )}
1880: $$
1881: \be
1882: \left. + {z^{-m} \over m!}~ {\pi ~\cos(m \pi/2)
1883: \over \sin((m+1)\pi/\mu) ~ \Gamma(\theta -(m+1)\theta/\mu )} \biggl]
1884: \right.
1885: \label{fox5} ~.
1886: \ee
1887: This expansion exists only for $\mu>\theta$ (conditions of page 71 below
1888: eq.~(3.7.2) of \cite{Mathai}). This is easily checked by the behavior
1889: of an asymptotics similar to (\ref{mgjgl}). Note that the series (\ref{fox5})
1890: is not defined in the
1891: special case $\mu=1$ due to the presence of the ill-defined ratio
1892: $\Gamma(0)/\Gamma(0)$ and a different approach is required, such as the
1893: integral representation of $W(t,\vec r)$ developed in \cite{Saichev}.
1894: The global Omori law obtained by integrating over the whole space
1895: (\ref{fox5})
1896: is again $N(t) \sim 1/t^{1-\theta}$ as expected from the analysis of
1897: the ETAS model
1898: without space dependence \cite{HS1}.
1899:
1900: Keeping only the largest term of (\ref{fox5}) for large $z$, we
1901: obtain the asymptotic
1902: behavior for small distances $ r < D~t^{\theta /\mu}$
1903: \ba
1904: N(t,r) &\simeq&
1905: {\Gamma(1-2\mu) ~
1906: ~\sin(\pi \mu) ~\sin(\pi \theta) ~ \over c' \sigma~ \pi^2}
1907: {\Gamma(1+\theta) \over (r/\sigma)^{1-2\mu}}~{1 \over (t/c')^{1+\theta}}
1908: ~~~\rm{for}~~ \mu<0.5 \nonumber \\
1909: N(t,r) &\simeq& { c'^{-\theta} \over
1910: c' \sigma~\mu~\Gamma(\theta-\theta/\mu) ~\sin(\pi/ \mu)}~
1911: {1 \over (t/c')^{1-\theta+\theta/\mu}}
1912: ~~~~~~~~~~~~~~~~\rm{for}~~ 0.5<\mu<2 ~.
1913: \label{lqps1}
1914: \ea
1915: Note that for $ r < D~t^{\theta /\mu}$ and $0.5<\mu<2$, the leading behavior
1916: of $N(t,r)$ is independent of $r$.
1917:
1918: Equation (\ref{lqps1}) thus predicts an apparent exponent
1919: \ba
1920: p &=& 1+\theta ~~~~~~~~~~~~~~~~\rm{for}~~ \mu<0.5 \nonumber \\
1921: p &=& 1-\theta + \theta/\mu ~~~~~\rm{for}~~ 0.5<\mu<2
1922: \label{leqps5}
1923: \ea
1924: for small distances $r < D~t^{\theta /\mu}$.
1925: This prediction is valid only in the case $\mu>\theta$ for which the series
1926: (\ref{fox5}) is convergent.
1927: However, the same asymptotic results are also obtained by different methods
1928: in the case $\mu<\theta$, for instance
1929: expression (\ref{leqps5}) is recovered for all $\mu<2$ using the
1930: integral representation of \cite{Saichev} [A. Saichev, private communication].
1931: The numerical evaluation of (\ref{njhaagrssjb}), which converges for
1932: $\mu < \theta$, also recovers the asymptotic results (\ref{lqps1}).
1933: The two regimes $\mu<0.5$ and $0.5<\mu<2$ are illustrated in Figures
1934: \ref{piuonk}
1935: and \ref{piuon} respectively. The seismicity rate $N(t,\vec r)$ is evaluated
1936: from expression (\ref{njhaagrssjb}) for small $z$ and from
1937: expression (\ref{fox5})
1938: for large $z$.
1939:
1940: We also performed numerical
1941: simulations of the ETAS and CTRW models and the results are in good
1942: agreement with
1943: expression (\ref{njhaagrssjb}) and (\ref{fox5}) for $N(\vec r,t)$ for $t \gg c$
1944: and $r \gg d$. For very small times $t\ll c$, or for very small distances
1945: $r \ll d$, expressions (\ref{njhaagrssjb}) and (\ref{fox5}) are not
1946: valid because
1947: they are based on a long wavelength $|\vec k| \to 0$ and long time
1948: $\beta \to 0$
1949: approximation.
1950: Numerical simulations of the ETAS model in the case $\theta=0.2$ and $\mu=0.9$
1951: are presented in Figure \ref{poiq}, and are in good agreement with
1952: the analytical
1953: solutions (\ref{njhaagrssjb}) and (\ref{fox5}) shown in Figure \ref{piuon}
1954: for the same parameters, except from the truncation of $N(t,r)$ for times
1955: $t\ll c$ and distances $r \ll d$ that are not reproduced by the
1956: analytical solution.
1957:
1958: \subsection{A simple non-separable joint distribution
1959: of waiting times and jump sizes: coupled spatial diffusion and long
1960: waiting time distribution}
1961:
1962: Consider the choice for $\phi_{m_i}(t-t_i, \vec r-\vec r_i)$
1963: replacing (\ref{first})
1964: by
1965: \be
1966: \phi_{m_i}(t-t_i, \vec r-\vec r_i) = \rho(m_i)~\Psi(t-t_i)~\Phi(|\vec
1967: r-\vec r_i|/\sqrt{D t})~,
1968: \label{firsaat}
1969: \ee
1970: where $\rho(m_i)$ and $\Psi(t)$ are again given by (\ref{formrho})
1971: and (\ref{psidef})
1972: while (\ref{phidef}) is changed into
1973: \be
1974: \Phi(|\vec r-\vec r_i|/\sqrt{D t}) = {1 \over \sqrt{2Dt}} ~
1975: \exp \left(-|\vec r-\vec r_i|^2/D t\right)~.
1976: \label{mgmglw}
1977: \ee
1978: The spatial diffusion of seismic activity is now coupled to the
1979: waiting time distribution. Expression (\ref{mgmglw}) captures the
1980: effect that, in order for aftershocks to spread over large distances
1981: by the underlying physical process, they need time. In fact,
1982: returning to the discussion in the introduction on the various
1983: proposed mechanisms for aftershocks, expression (\ref{mgmglw})
1984: embodies a microscopic diffusion process.
1985:
1986: In this case, (\ref{ngnslqw}) must be replaced by
1987: \be
1988: {\hat N}(\beta,\vec k) = {{\hat S}_M(\beta,\vec k) \over
1989: 1 - n {\hat \phi}(\beta, \vec k)}~,
1990: \label{ngssnslqw}
1991: \ee
1992: where ${\hat \phi}(\beta, \vec k)$ is the Laplace-Fourier transform
1993: of the product
1994: $\Psi(t)~\Phi(|\vec r|/\sqrt{D t})$. For large times and long distances for
1995: which the first terms in the expansion in $\beta$ and $k$ are sufficient, and
1996: for $n=1$, we obtain
1997: \be
1998: {\hat \phi}(\beta, \vec k) \propto {{\hat S}_M(\beta,\vec k) \over
1999: (\beta + D k^2)^{\theta}}~.
2000: \label{gjjgws}
2001: \ee
2002: The inverse Laplace-Fourier transform of (\ref{ngssnslqw}) is
2003: \be
2004: N(t, \vec r) \sim {1 \over t^{1-\theta}}~{1 \over \sqrt{2 \pi Dt}}~
2005: \exp\left(-|\vec r|^2/D t\right)~.
2006: \label{mjgwlw}
2007: \ee
2008: As expected, expression (\ref{mjgwlw}) recovers the dressed Omori propagator
2009: in the case of absence of space dependence \cite{HS1}. At finite $r$
2010: and long times,
2011: the dressed Omori law also decay as $1 /t^{1-\theta}$. The diffusion of
2012: aftershocks is normal with the standard diffusion exponent $H=1/2$.
2013:
2014:
2015:
2016: \section{New questions on aftershocks derived from the CTRW analogy}
2017:
2018: We list a series of comments and questions suggested from the analogy
2019: between the ETAS model and the CTRW model. In particular, we discuss the
2020: possibility of defining new observables for earthquake aftershocks, that could
2021: be worthwhile to investigate in future empirical studies of
2022: earthquake aftershocks.
2023:
2024: \subsection{Recurrence of aftershock activity in the proximity of the
2025: main shock}
2026:
2027: A quantity often investigated in studies of random walks is the
2028: probability $W(t, \vec 0)$
2029: to find the random walker at its starting point (the origin) at time $t$.
2030: In the earthquake framework, this is the seismic aftershock rate
2031: close to the main shock.
2032:
2033: \subsection{First-passage times}
2034:
2035: The first passage time of a random walk is the first arrival time
2036: of the random walk at a given point $\vec r$. In the earthquake context,
2037: this translates into the study of the waiting time for a given region
2038: to have its own first aftershock after the main shock occurs.
2039: The distribution of such first passage waiting times gives the distribution
2040: of times with no nearby seismic activity. See for instance \cite{barkai1}
2041: in the case of a power law distribution of waiting times and Gaussian
2042: distribution of jump sizes. Margolin and Berkowitz \cite{Margolin} give
2043: the distribution of first-passage times in the case where the jump distribution
2044: is narrow and the waiting distribution is long-tailed $\sim 1/t^{1+\theta}$.
2045: They analyze the three different regimes $\theta <1$, $1<\theta <2$
2046: and $\theta \geq 2$.
2047:
2048:
2049: \subsection{Occupation time of seismic activity}
2050:
2051: Weiss and Calabrese \cite{Weiss2} have studied the total amount of time spent
2052: by a lattice CTRW on a subset of points. In the seismic language,
2053: this amounts to study the probability distribution of the durations
2054: of aftershock sequences that are localized in a specific subset of the space.
2055: In other words, how probable are aftershock sequences that are found only
2056: within a given spatial subset over a certain duration?
2057:
2058:
2059: \subsection{Transience and recurrence of seismic activity}
2060:
2061: Another question that has been studied in some details in the CTRW framework
2062: is whether random walks are transient or recurrent. A transient random walk
2063: visits any point $\vec r$ at most a finite number of times before escaping to
2064: infinity. For earthquakes, the transient regime corresponds to the activation
2065: of at most a finite number of aftershocks in any given point $\vec r$.
2066: In contrast, a recurrent random walk may return a growing number of times
2067: to all or a subset of points at time increases. In the aftershock language,
2068: this means that these points will have a
2069: never-ending (decaying) aftershock activity. We stress here the difference
2070: between the global Omori law giving a never-ending power law decay of
2071: the aftershock activity (in the sub-critical regime $n<1$) and its
2072: spatial dependence which must exhibit important variations. In particular,
2073: in the recurrent regime, an Omori law can be documented by counting
2074: aftershocks in those limited regions of space which are activated
2075: again and again.
2076:
2077: \subsection{Probability for the cumulative number of aftershocks}
2078:
2079: Let us define a basic quantity in the CTRW formalism, namely the probability
2080: $\chi_m(t)$ to make exactly $m$ steps up to time $t$. In the
2081: earthquake context,
2082: $\chi_m(t)$ is the probability to have exactly $m$ aftershocks after
2083: the main shock.
2084: In the case in which the spatial transition probability $\Phi(\vec r)$ between
2085: different positions is independent of the waiting times (corresponding to
2086: factorizing $\phi_{m_i}(t-t_i, \vec r-\vec r_i)$ as in (\ref{first})), the
2087: probability density $W(t, \vec r)$ to find the walker at position $\vec r$ at
2088: time $t$ can be written
2089: \be
2090: W(t, \vec r) = \sum_{m=0}^{+\infty} W_m(\vec r)~\chi_m(t)~,
2091: \label{gkmkvms}
2092: \ee
2093: where $W_m(\vec r)$ is the probability to reach $\vec r$ from $\vec
2094: 0$ in $m$ steps.
2095: In the earthquake context, $W_m(\vec r)$ is the probability that
2096: there has been exactly
2097: $m$ events in the time interval $[0,t]$ and that the last one
2098: occurred at $\vec r$.
2099: Equation (\ref{gkmkvms}) states that the CTRW is a random process
2100: subordinated to
2101: simple random walks described by $W_m(\vec r)$
2102: under the operational time given by the $\chi_m(t)$ distribution
2103: \cite{Sokolov1,Sokolov2}.
2104:
2105:
2106: \subsection{Random walk models with birth and death and background
2107: seismicity from localized sources}
2108:
2109: Bender et al. \cite{Bender} have studied models of random walks in
2110: which walkers are born
2111: in proportion to the population
2112: at one specific site (for instance the origin)
2113: with probability $a-1$ (with $a>1$) and die at all
2114: other sites with probability $1-n$ (with $n \leq 1$). In the
2115: earthquake context, this consists in assuming that the aftershock activity is
2116: fed by a localized region in space, which is itself activated by the aftershocks
2117: returning to this region, furthering the overall activity. This may
2118: be considered
2119: to describe the seismic activity close to a plate boundary, in which the plate
2120: boundary is the constant self-consistent source of a seismic activity
2121: which may spread over
2122: a significant region away from the boundary. The excursion of the
2123: random walkers
2124: quantify the spread of the seismic activity away from the main fault structure.
2125: The rate of death of the walkers correspond exactly to the distance $1-n$ from
2126: the critical value $n=1$. Bender et al. \cite{Bender} find a phase
2127: diagram in the
2128: $(a-1, 1-n)$ parameter space in which a boundary separates two possible
2129: asymptotic regimes:
2130: \begin{enumerate}
2131: \item for small $a-1$ and large $1-n$, the seismic activity at the origin
2132: and everywhere eventually dies off;
2133: \item for large $a-1$ and small $1-n$, the average seismic activity
2134: at the origin approaches
2135: a positive constant at long times. In this regime, there is a
2136: transition as $a-1$
2137: is decreased or as $1-n$ is increased, between
2138: a case where the global seismic activity outside the origin goes to
2139: zero and a case where it
2140: diverges at long times. On the boundary between these two regimes in the
2141: $(a-1, 1-n)$ parameter space, the distribution of seismic activity
2142: approaches a steady state
2143: at long times. There is a critical point (for space dimensions
2144: different from $2$)
2145: at a certain value $(a_c-1, 1-n_c)$, for
2146: which the long-time seismic activity away from the source is given by
2147: $\sim (a - a_c)^{\nu}$
2148: where $\nu$ is a critical exponent equal to $2$ in three dimensions.
2149: \end{enumerate}
2150:
2151: Note that the results of \cite{Bender} are obtained for random
2152: walks on a lattice.
2153: This can easily be converted into a CTRW by the fact that
2154: a CTRW is nothing by a process subordinated to
2155: discrete random walks under the operational time defined by the
2156: process $\{t_i\}$
2157: of the time of just arrival to a given site, as given by (\ref{gkmkvms}).
2158:
2159:
2160: \section{Discussion}
2161:
2162: Using the analogy between the ETAS model and the CTRW model
2163: established here, we have derived the relation between the average
2164: distance between aftershocks and the mainshock as a function of the
2165: time from the mainshock, and the joint probability distribution
2166: of the times and locations of aftershocks.
2167:
2168: We have assumed that each earthquake triggers aftershocks at a distance $r$
2169: and time $t$ according to the bare propagator $\phi(r,t)$, which can
2170: be factorized as $\Psi(t)\Phi(r)$. This means that the distribution
2171: $\Phi(r)$ of the distances between an event and its direct aftershocks is
2172: decoupled from the distribution $\Psi(t)$ of waiting time. Hence, the direct
2173: aftershocks triggered by a single mainshock do not diffuse in space with time.
2174: Notwithstanding this decoupling in space and time of the bare propagator
2175: $\phi(r,t)$, we have shown that the global law or dressed propagator
2176: $N(t,\vec r)$ defined as the global rate of events at time $t$ and at position $\vec r$,
2177: cannot be factorized into two distributions of waiting times and space jumps.
2178: This joint distribution of waiting times and positions of the whole sequence of
2179: aftershocks cascading from a mainshock is different from the product
2180: of the bare time and space propagators.
2181:
2182: The mean distance between the mainshock and its aftershocks, including
2183: secondary aftershocks, increases with the time from the mainshock,
2184: due to the cascade process of aftershocks triggering aftershocks triggering
2185: aftershocks, and so on.
2186: In the critical case $n=1$, this diffusion takes the form
2187: of a power-law relation $R \sim t^H$ of the average distance $R$ between
2188: aftershocks and the mainshock, as a function of the time $t$ from the
2189: mainshock.
2190: If the local Omori law is characterized by an exponent $0<\theta<1$, and if
2191: the space jumps follow a power law $\Phi(r) \sim 1/(r +d )^{1+\mu}$,
2192: the diffusion exponent is given by $H=\theta/\mu$ in the case $\mu<2$ and
2193: $H=\theta/2$ in the case $\mu>2$. Depending on the $\theta$ and $\mu$ values,
2194: we can thus observe either sub-diffusion ($H<1/2$) or super-diffusion
2195: ($H>1/2$), as summarized in Figure \ref{H}.
2196: In the sub-critical ($n<1$) and super-critical ($n>1$) regimes,
2197: this relation is still valid up to the characteristic time $t^*$
2198: given by (\ref{tstar})
2199: and for distances smaller than $r^* \propto D t^{*H}$ given by (\ref{gnjgrkd}).
2200: For $t>t^*$ and $r>r^*$ in the sub-critical regime,
2201: the global distributions of times and distances
2202: between the mainshock and its aftershocks are decoupled and there is therefore
2203: no diffusion. In the super-critical regime, the aftershock rate increases
2204: exponentially for $t>t^*$ and the aftershocks diffuses more rapidly
2205: than before $t^*$.
2206:
2207:
2208: In the critical regime, the cascade of secondary aftershocks introduces a
2209: variation of the apparent Omori exponent as a function of the
2210: distance from the mainshock. The asymptotic values of the Omori exponent
2211: in the different regimes are summarized in Table \ref{table2}.
2212: In the regime $\mu<2$, we observe a transition from an Omori law decay
2213: with an exponent $p=1-2\theta$ at early times $t^H \ll r/D$ to
2214: a larger exponent at large times. This provides
2215: another mechanism to explain the observed variability of the Omori exponent.
2216: In the regime $\mu>2$, a power-law decay of the seismicity with time is
2217: observed only at large times $t^H \gg r/D$. At early times, or at large
2218: distances $r \gg D t^H$, the seismicity rate is very small, because the
2219: seismicity as not yet diffused up to the distance $r$.
2220:
2221: We should emphasize that our theoretical analysis of aftershock
2222: diffusion predicts the behavior of the ensemble average of aftershock sequences.
2223: Individual
2224: sequences may depart from this ensemble average, especially for
2225: sequences with few earthquakes and limited durations. For long sequences
2226: (20,000 events say), we have verified that the exponent $H$
2227: measured on individual sequences does not deviate from the ensemble
2228: average value by more than about 20\%. As already discussed, the
2229: impact of fluctuations become however more effective as the parameter $\alpha$
2230: increases above $b/2$.
2231:
2232: The diffusion of the seismicity also renormalizes the spatial distribution
2233: of the seismicity, which is very different from the local
2234: distribution $\Phi(r)$
2235: of distances between a triggering event and its direct aftershocks.
2236: In the regime $\mu>2$, the global seismicity rate $N(t,\vec r)$ decays
2237: exponentially with the distance from the mainshock, whereas the local
2238: distribution of distances $\Phi(r)$ is a power-law distribution.
2239: In the regime $\mu<2$, the local law $\Phi(r) \sim r^{-1-\mu}$ is
2240: recovered at large distances, but a slower decay for $0.5<\mu<2$ or a constant
2241: rate for $\mu<0.5$ is observed at small distances $r \ll D t^H$.
2242: These predictions on the decrease of the Omori exponent with $r$
2243: have not yet been observed in earthquake catalogs, but
2244: an expansion of the aftershock zone has been reported in many studies
2245: \cite{Mogi,Imoto,Chatelain,Tajima1,Tajima2,Wesson,Ouchi,Noir,Jacques}.
2246: However, very few studies
2247: have quantified the diffusion law. Noir et al. \cite{Noir}
2248: show that the earthquake Dobi sequence (central Afar, August 1989)
2249: composed of 22 $M>4.6$ earthquakes presented a migration that was
2250: in agreement with a diffusion process due to fluid transfer in the crust,
2251: characterized by a normal diffusion process with exponent $H=0.5$.
2252: Tajima and Kanamori \cite{Tajima1,Tajima2} studied several aftershock
2253: sequences in subduction zone and observed a much slower logarithmic
2254: diffusion, which is compatible with a low exponent $H$ close to $0.1$.
2255: In some cases, the aftershock sequence displays no expansion with time.
2256: For instance, Shaw \cite{Shaw} studied several aftershock sequences in
2257: California and concluded that the distribution of distances between the
2258: mainshock and its aftershocks is independent of time.
2259: This can be explained by the fact that the 0mori exponent measured
2260: in \cite{Shaw} is very close to $1$, thus $\theta$ is very small and our
2261: prediction is that the exponent $H$ should be very small.
2262:
2263: In fact, the ETAS model predicts that diffusion should be observed only
2264: for aftershock sequences with a measured Omori exponent $p$ significantly
2265: smaller than $1$, which can only occur according to our model when the bare
2266: Omori propagator with exponent $1+\theta$ is renormalized into the
2267: dressed propagator with global exponent $1-\theta$. We have shown that
2268: this renormalization of the exponent only occurs at times less than $t^*$,
2269: while for longer times in the sub-critical regime $n<1$ the dressed Omori
2270: propagator recovers the value of the bare exponent $1+\theta > 1$
2271: (see figure \ref{n}).
2272: Therefore, identifying an empirical observation of $p<1$ with our
2273: prediction $p=1-\theta$ indicates that the aftershock sequence falls in the
2274: ``good'' time window $t<t^*$ in which the renormalization operates.
2275: We have also shown that the dressed propagator gives a diffusion only for
2276: $t<t^*$. We can thus conclude that, according to the ETAS model, the
2277: observation of an empirical Omori
2278: exponent larger than $1$ is indicative of the large time $t>t^*$ behavior
2279: in the sub-critical regime $n<1$, for which there is no diffusion.
2280: This provides a possible explanation for why many sequences studied by
2281: \cite{Tajima1,Tajima2,Shaw} do not show a diffusion of the aftershock
2282: epicenters.
2283: Reciprocally, a prerequisite for observing diffusion in
2284: a given aftershock sequence is that the empirical
2285: $p$-value be less than $1$ in order to qualify the regime $t<t^*$.
2286:
2287: An alternative model has been
2288: discussed by Dieterich \cite{Diete} who showed that the spatial
2289: variability of the stress
2290: induced by a mainshock, coupled with a rate and state friction law, results
2291: in an expansion of the aftershock zone with time. This expansion does not take
2292: the form of a diffusion law as observed in the ETAS model,
2293: the relation between the characteristic size
2294: of the aftershock zone does not grow as a power law of the time from the
2295: mainshock (equation (22) and Figure 6 of \cite{Diete}).
2296:
2297: Marsan et al. \cite{Marsan1,Marsan2} and Marsan and Bean \cite{Marsan3}
2298: studied several catalogs at different scales, from the scale of a
2299: deep mine to the world-wide seismicity, and observed that the average
2300: distance between two earthquakes increases as a power-law of the time between
2301: them, with an exponent often close to 0.2, indicative of a sub-diffusion process.
2302: They interpreted their results as a mechanism of stress diffusion, that may
2303: be due to fluid transfer with heterogeneous permeability leading to sub-diffusion.
2304: Their analysis is quite different from those used in other studies, because
2305: they consider all pairs of events, without distinction between
2306: aftershocks and mainshocks.
2307: This analysis can however lead to spurious diffusion, and in some
2308: cases this method does not detect diffusion in synthetic data set
2309: with genuine diffusion.
2310: We have tested their analysis on a synthetic catalog generated by superposing
2311: a background seismicity with uniform spatial and temporal distribution, and
2312: 10 mainshocks with poissonian distribution in time and space, and with a
2313: power-law distribution of energies. Each of these mainshocks generates only
2314: {\it direct} aftershocks, without secondary cascades of aftershocks,
2315: and the number of aftershocks increases exponentially with the
2316: magnitude of the mainshock.
2317: This way, we generate a synthetic catalog without any physical
2318: process of diffusion, and which includes all the other well-established
2319: characteristics of real seismicity: clustering in space and time superposed
2320: to a seismicity background.
2321: Applying the analysis of \cite{Marsan1,Marsan2,Marsan3} to this
2322: synthetic data set leads to an apparent
2323: diffusion process with a well-defined exponent $H=0.5$.
2324: However, this apparent diffusion does not
2325: reflect a genuine diffusion but simply describes the crossover
2326: from the characteristic size of an aftershock zone at early times to the
2327: larger average distance between uncorrelated events at large times. In plain
2328: words, the apparent power law $R \propto t^H$ is nothing but a cross-over
2329: and is not real.
2330: Furthermore, applying this analysis to a synthetic catalog generated
2331: using the ETAS model, without seismicity background, and with a theoretical
2332: diffusion exponent $H=0.2$, the method yields $H=0.01$ if we use all the
2333: events of the catalog. If we select only events up to a maximum
2334: distance $r_{max}$ to apply the same procedure as in
2335: \cite{Marsan1,Marsan2,Marsan3},
2336: we obtain larger values of $H$ which are more in agreement with the
2337: theoretical exponent $H=0.2$ but with large fluctuations that are function of
2338: $r_{max}$.
2339: Therefore, it is probable
2340: that the diffusion reported in \cite{Marsan1,Marsan2,Marsan3} is not real
2341: and results from a cross-over between two characteristic
2342: scales of the spatial earthquake distribution. It may be attributed to
2343: the analyzing methodology which mixes up uncorrelated events.
2344: We are thus reluctant to compare the results of Marsan et al.
2345: \cite{Marsan1,Marsan2,Marsan3}
2346: with the predictions obtained with the ETAS model.
2347:
2348: One can similarly question the results on anomalous diffusion of seismicity
2349: obtained by Sotolongo-Costa et al. \cite{Sotolongo}, who
2350: considered 7500 micro-earthquakes recorded by a local spanish network
2351: from 1985 to 1995. They interpret the sequence of earthquakes
2352: as a random walk process, in which the walker jumps from an
2353: earthquake epicenter
2354: to the next in sequential order. The time between two successive events is
2355: seen as a waiting time between two jumps and the distance between these
2356: events is taken to correspond to the jump size. Since the distributions of
2357: time intervals and of distances between successive earthquakes are both
2358: heavy-tailed (approximately power laws), their model is a CTRW.
2359: We cannot stress enough that their CTRW model of seismicity has nothing to
2360: do with our results on the mapping of the ETAS model onto a CTRW.
2361: Their procedure is ad-hoc and their results depend obviously strongly on
2362: the space domain of the analysis since distant earthquakes that are
2363: completely unrelated can be almost simultaneous!
2364: We also stress that our mapping of the ETAS model onto the CTRW model
2365: does not correspond to identifying an earthquake sequence as a {\it single}
2366: realization of a CTRW, as assumed arbitrarily by Sotolongo-Costa et al.
2367: \cite{Sotolongo}.
2368:
2369: Our predictions obtained here are thus difficult to test on
2370: seismicity data, due to the small
2371: number of events available and the restricted time periods and
2372: distance ranges,
2373: and because the seismicity background can strongly affect the results.
2374: New methods should hence be developed to investigate if there is a
2375: real physical
2376: process of diffusion in seismic activity and to
2377: compare the observations of real seismicity with the quantitative predictions
2378: of the ETAS model.
2379: Preliminary study of aftershock sequences in California leads to the conclusion
2380: that most aftershock sequences are characterized by an Omori exponent $p>1$,
2381: indicative of the sub-critical regime with $t>t^*$. As expected from our
2382: predictions in this regime, we do not observe
2383: an expansion of the aftershock zone.
2384: However, a few sequences give a value $p<1$ and also exhibit an
2385: increase of the average
2386: distance between the mainshock and its aftershocks consistent with our
2387: predictions. A detailed report of this analysis will be reported elsewhere.
2388:
2389:
2390: \section{Conclusion}
2391:
2392: We have studied analytically and numerically the ETAS
2393: (epidemic-type aftershock) model, which is a simple
2394: stochastic process modeling seismicity, based on the two best-established
2395: empirical laws for earthquakes, the power law decay of seismicity
2396: after an earthquake and a power law distribution of earthquake energies.
2397: This model assumes that each earthquake can trigger aftershocks, with
2398: a rate increasing with its magnitude.
2399: In this model, the seismicity rate is the result of the whole cascade of
2400: direct and secondary aftershocks.
2401:
2402: We have first established an exact correspondence between the ETAS model
2403: and the CTRW (continuous-time random walk) model. We have then used
2404: this analogy to derive the joint probability of times and distances of the
2405: seismicity following a large earthquake and we have characterized
2406: the different regimes of diffusion.
2407:
2408: We have shown that the diffusion of the seismicity should be observed only
2409: for times $t<t^*$, where $t^*$ is a characteristic time depending on the model
2410: parameters, corresponding to an observed Omori exponent smaller than one.
2411: Most aftershock sequences have an observed Omori exponent larger than one,
2412: corresponding to the subcritical regime of the ETAS model, for which there
2413: is no diffusion.
2414: The diffusion of the seismicity produces a decrease of the Omori
2415: exponent as a function of the distance from the mainshock, the decay
2416: of aftershocks being faster close to the mainshock than at large distances.
2417: The spatial distribution of seismicity is also renormalized by the
2418: cascade process,
2419: so that the observed distribution of distances between the mainshock
2420: and its aftershocks can be fundamentally different from the bare
2421: propagator $\Phi(r)$ which gives the distribution of the distances
2422: between triggered and triggering earthquakes. We have also
2423: noted that the ETAS model generates apparent but realistical fractal
2424: spatial patterns.
2425:
2426: Assuming that the distances between triggering and triggered events
2427: are independent of the time between them, this model generates a diffusion of
2428: the whole sequence of aftershocks with the time from the mainshock,
2429: which is induced by the cascade of aftershocks triggering aftershocks, and so on.
2430: Our results thus provides a simple explanation of the diffusion of
2431: aftershock sequences
2432: reported by several studies, which was often interpreted as a mechanism
2433: of anomalous stress diffusion. We see that no such ``anomalous stress
2434: diffusion'' is
2435: needed and our theory provides a parsimonious account of aftershock diffusion
2436: resulting from the minimum physical ingredients of the ETAS model.
2437: As Einstein once said, ``A theory is more impressive the greater
2438: the simplicity of its premises, the more different the kinds of things
2439: it relates and the more extended its range of applicability.''
2440:
2441: \acknowledgments
2442: We are very grateful to B. Berkowitz, S. Gluzman, J.-R. Grasso, Y. Klafter, L. Margerin,
2443: A. Saichev and G. Zaslavsky for useful suggestions and discussions.
2444:
2445:
2446:
2447: \begin{thebibliography}{}
2448:
2449: \bibitem{Geilo} Sornette, D.,
2450: Spontaneous formation of space-time structures and criticality,
2451: Vol. 349 of NATO Advance Study Institute; Series B: Physics,
2452: edited by T. Roste and D. Sherrington, (Kluwer, Dordrecht; Boston, 1991),
2453: pp.57-106.
2454:
2455: \bibitem{Rundleklein} Rundle, J.B. and W. Klein, New ideas about the
2456: physics of earthquakes, Reviews of Geophysics 33, 283-286, 1995.
2457:
2458: \bibitem{Main} Main, I., Statistical physics, seismogenesis and seismic
2459: hazard, Review of Geophysics 34, 433-462, 1996.
2460:
2461: \bibitem{Sorreview} Sornette, D.,
2462: Earthquakes: from chemical alteration to mechanical rupture, Physics
2463: Reports 313, 238-291, 1999.
2464:
2465: \bibitem{Turcottereview} Turcotte, D.L.,
2466: Self-organized criticality, Reports on Progress in Physics 62, 1377-1429, 1999.
2467:
2468: \bibitem{GR} Gutenberg, B. and C.F. Richter, Frequency of earthquakes
2469: in California,
2470: Bull. Seism. Soc. Am. 34, 185-188, 1944.
2471:
2472: \bibitem{Omori} Omori, F., On the aftershocks of earthquakes,
2473: J. Coll. Sci. Imp. Univ. Tokyo 7, 111-120, 1894.
2474:
2475: \bibitem{Ouillonetal} Ouillon, G., C. Castaing and D. Sornette,
2476: Hierarchical scaling of faulting, J. Geophys. Res. 101 B3, 5477-5487, 1996.
2477:
2478: \bibitem{KK80} Kagan Y.Y. and L. Knopoff,
2479: Spatial distribution of earthquakes: the two-point correlation
2480: function, Geophys. J. R. astr. Soc. 62, 303-320, 1980.
2481:
2482:
2483:
2484: \bibitem{Pissor} Pisarenko, V.F. and D. Sornette,
2485: Characterization of the frequency of extreme events by the Generalized
2486: Pareto Distribution, in press in Pure and Applied Geophysics, 2002
2487: (e-print cond-mat/0011168)
2488:
2489:
2490: \bibitem{kaganuniv} Kagan, Y.Y.,
2491: Universality of the seismic moment-frequency relation,
2492: Pure and Applied Geophysics, 155, 537-573, 1998.
2493:
2494: \bibitem{Sorbook} Sornette, D., Critical Phenomena in Natural Sciences,
2495: Springer Series in Synergetics, Heidelberg, 2000.
2496:
2497: \bibitem{Harrisreview} Harris, R.A., Stress triggers, stress shadows,
2498: and seismic hazard, International Handbook of Earthquake and
2499: Engineering Seismology,
2500: W.H.K. Lee, H. Kanamori, P. C. Jennings, C. Kisslinger, Eds., Chapter
2501: 73, 48 pages, in press.
2502:
2503: \bibitem{Yamakno} Yamashita, T. and Knopoff, L.,
2504: Models of aftershock occurrence, Geophys. J. Royal Astron. Soc. 91,
2505: 13-26, 1987.
2506:
2507: \bibitem{Leesor} Lee, M.W. and D. Sornette,
2508: Novel mechanism for discrete scale invariance in sandpile models,
2509: European Physical Journal B 15, 193-197, 2000.
2510:
2511:
2512: \bibitem{Huangetal} Huang, Y., H. Saleur, C. G. Sammis, D. Sornette,
2513: Precursors, aftershocks, criticality and self-organized criticality,
2514: Europhysics Letters 41, 43-48, 1998.
2515:
2516: \bibitem{Utsu} Utsu, T., Y. Ogata and S. Matsu'ura, The centenary of the
2517: Omori Formula for a decay law of aftershock activity,
2518: J. Phys. Earth 43, 1-33, 1995.
2519:
2520: \bibitem{BathRichter} Bath, M. and C.F. Richter,
2521: Mechanisms of the aftershocks of
2522: the Kern County, California earthquake of 1952, Bull. Seism. Soc. Am.
2523: 48, 133-146, 1958.
2524:
2525: \bibitem{Beroza} Beroza, G.C. and M.D. Zoback, Mechanism diversity
2526: of the Loma-Prieta aftershocks and the mechanics of mainshock-aftershock
2527: interaction, Science 259, 210-213, 1993.
2528:
2529: \bibitem{Hill} Hill, D.P. et al., Seismicity remotely triggered by
2530: the magnitude 7.3
2531: Landers, California, earthquake, Science 260, 1617-23, 1993.
2532:
2533: \bibitem{Steeples} Steeples, D.W. and D.D. Steeples, Far-field
2534: aftershocks of the 1906 earthquake, Bull. Seism. Soc. Am. 86, 921-924, 1996.
2535:
2536:
2537: \bibitem{KJ1} Kagan, Y.Y. and D.D. Jackson,
2538: Spatial aftershock distribution: Effect of normal stress, J.
2539: Geophys. Res. 103, 24453-24467, 1998.
2540:
2541: \bibitem{Meltzner1} Meltzner, A.J. and D.J. Wald,
2542: Foreshocks and aftershocks of the great 1857 California earthquake,
2543: Bull. Seism. Soc. Am. 89, 1109-1120, 1999.
2544:
2545: \bibitem{Dreger} Dreger, D. and B. Savage,
2546: Aftershocks of the 1952 Kern County, California, earthquake sequence,
2547: Bull. Seism. Soc. Am. 89, 1094-1108, 1999.
2548:
2549: \bibitem{SS} Sornette, A. and D. Sornette, Renormalization of
2550: earthquake aftershocks,
2551: Geophys. Res. Lett. 26, 1981-1984, 1999.
2552:
2553: \bibitem{HS1} Helmstetter, A. and D. Sornette,
2554: Sub-critical and Super-critical Regimes in Epidemic Models of
2555: Earthquake Aftershocks, J. Geophys. Res, in press, 2002.
2556: (http://arXiv.org/abs/cond-mat/0109318)
2557:
2558: \bibitem{SH2} Sornette, D. and A. Helmstetter,
2559: New Mechanism for Finite-Time-Singularity in Epidemic Models of Rupture,
2560: Earthquakes and Starquakes, Phys. Rev. Lett., 89, 158501 (2002).
2561: (e-print at cond-mat/0112043).
2562:
2563: \bibitem{Mogi} Mogi K., Sequential occurrences of recent great earthquakes,
2564: J. Phys. Earth 16, 30, 1968.
2565:
2566: \bibitem{Imoto} Imoto,M., On migration phenomena of aftershocks
2567: following large thrust earthquakes in subduction zones,
2568: Report of the National Research Center for Disaster Prevention,
2569: 25, 29-71, 1981.
2570:
2571: \bibitem{Chatelain} Chatelain, J.L., R.K. Cardwell and B.L.Isacks,
2572: Expansion of the aftershock zone following the Vanuatu (New Hebrides)
2573: earthquake on 15 July 1981, Geophys. Res. Lett. 10, 385-388, 1983.
2574:
2575: \bibitem{Tajima1} Tajima, F. and H. Kanamori, Global survey of aftershock area
2576: expansion patterns, Phys. Earth Planet. Inter. 40, 77-134, 1985a.
2577:
2578: \bibitem{Tajima2} Tajima, F. and H. Kanamori, Aftershock area
2579: expansion and mechanical
2580: heterogeneity of fault zone within subduction zone,
2581: Geophys. Res. Lett. 12, 345-348, 1985b.
2582:
2583: \bibitem{Wesson} Wesson, R. L., Modeling aftershock migration and afterslip
2584: of the San Juan Bautista, California, earthquake of October 3, 1972,
2585: Tectonophysics 144, 214-229, 1987.
2586:
2587:
2588: \bibitem{Ouchi} Ouchi, T. and T. Uekawa, Statistical analysis of the
2589: spatial distribution
2590: of earthquakes - variation of the spatial distribution of
2591: earthquakes before
2592: and after large earthquakes, Phys. Earth Planet. Inter. 44,
2593: 211-225, 1986.
2594:
2595: \bibitem{Noir} Noir, J., E. Jacques, S. Bekri, P.M. Adler, P. Tapponnier, and
2596: G.C.P. King, Fluid flow triggered migration of events in the 1989 Dobi
2597: earthquake sequence of Central Afar.
2598: Geophys. Res. Lett. 24, 2335-2338, 1997.
2599:
2600: \bibitem{Jacques} Jacques, E, J.C. Ruegg, J.C. Lepine, P. Tapponnier,
2601: G.C.P. King, GCPand A. Omar, Relocation of $M \geq 2$ events of
2602: the 1989 Dobi seismic sequence in Afar: evidence for earthquake migration.
2603: Geophys. J. Int. 138, 447-469, 1999.
2604:
2605: \bibitem{Rydelek} Rydelek, P.A. and I.S. Sacks, Migration of large earthquakes
2606: along the San Jacinto fault; stress diffusion from 1857 Fort Tejon
2607: earthquake, Geophys. Res. Lett. 28, 3079-3082, 2001.
2608:
2609:
2610: \bibitem{Nur} Nur A. and J.R. Booker, Aftershocks caused by pore fluid flow ?
2611: Science 175, 885-888, 1972.
2612:
2613: \bibitem{Hudnut} Hudnut K.W., L. Seeber and J. Pacheco, Cross-fault
2614: triggering in the
2615: November 1987 Superstition Hill earthquake sequence, southern
2616: California, Geophys. Res. Lett. 16, 199-203, 1989.
2617:
2618: \bibitem{Diete} Dieterich, J., Constitutive law for rate of
2619: earthquake production and
2620: its application to earthquake clustering, J. Geophys. Res. 99,
2621: 2601-2618, 1994.
2622:
2623: \bibitem{KK} Kagan, Y.Y. and L. Knopoff,
2624: Statistical short-term earthquake prediction,
2625: Science 236, 1563-1467, 1987.
2626:
2627: \bibitem{Ogata1} Ogata, Y., Statistical models for earthquake occurrence
2628: and residual analysis for point processes, J. Am. stat.
2629: Assoc. 83, 9-27, 1988.
2630:
2631:
2632: \bibitem{YaShi} Yamanaka, Y. and K. Shimazaki, Scaling relationship between
2633: the number of aftershocks and the size of the mainshock,
2634: J. Phys. Earth 38, 305-324, 1990.
2635:
2636:
2637: \bibitem{Drakatos} Drakatos, G. and J. Latoussakis,
2638: A catalog of aftershock sequences in Greece (1971, 1997):
2639: their spatial and temporal characteristics,
2640: Journal of Seismology 5, 137 (2001).
2641:
2642: \bibitem{Guo2} Guo, Z. and Y. Ogata, Statistical relations between the
2643: parameters of aftershocks in time, space and magnitude,
2644: J. Geophys. Res. 102, 2857-2873, 1997.
2645:
2646: \bibitem{H02} Helmstetter, A. Is stress triggering driven by small
2647: earthquakes? submitted to Geophys. Res. Lett., 2002.
2648:
2649:
2650: \bibitem{Felzer} Felzer, K.R., T.W. Becker, R.E. Abercrombie, G. Ekstr\"om and
2651: J.R. Rice, Triggering of the 1999 $M_W$ $7.1$ Hector Mine earthquake by
2652: aftershocks of the 1992 $M_W$ $7.3$ Landers earthquake,
2653: submitted to the J. Geophys. Res. (Solid Earth)
2654:
2655: \bibitem{Bath} Bath, M., Lateral inhomogeneities in the upper mantel,
2656: Tectonophysics 2, 483-514, 1965.
2657:
2658: \bibitem{consolebath} Console, R., A.M. Lombardi, M. Murru and D. Rhoades,
2659: Bath's law and the self-similarity of earthquakes, in press
2660: in J. Geophys. Res., 2002.
2661:
2662: \bibitem{Ogata2} Ogata, Y., Statistical model for standard seismicity and
2663: detection of anomalies by residual analysis, Tectonophysics
2664: 169, 159-174, 1989.
2665:
2666: \bibitem{Ogata3} Ogata, Y., Detection of precursory relative quiescence before
2667: great earthquakes though a statistical model, J. Geophys. Res.
2668: 97, 19845-19871, 1992.
2669:
2670: \bibitem{Ogata5} Ogata, Y., Seismicity analysis through point-process
2671: modeling : a review, Pure Appl. Geophys. 155, 471-507, 1999.
2672:
2673: \bibitem{Ogata6} Ogata, Y., Increased probability of large earthquakes near
2674: aftershock regions with relative quiescence, J. Geophys. Res.
2675: 106, 8729-8744, 2001.
2676:
2677: \bibitem{Kagan} Kagan, Y.Y., Likelihood analysis of earthquake catalogues
2678: Geophys. J. Int. 106, 135-148, 1991
2679:
2680:
2681: \bibitem{KJ2} Kagan, Y.Y. and D.D. Jackson, Probabilistic forecasting of
2682: earthquakes, Geophys. J. Int. 143, 438-453, 2000.
2683:
2684: \bibitem{Console} Console, R. and M. Murru, A simple and testable
2685: model for earthquake clustering, J. Geophys. Res. 166, 8699-8711, 2001.
2686:
2687: \bibitem{Montroll1} Montroll, E.W. and G.H. Weiss, J. Math. Phys., 6,
2688: 167, 1965.
2689:
2690: \bibitem{Tanguy} Krishnamurthy, S., Tanguy, A., Abry, P. and Roux, S.,
2691: A stochastic description of extremal dynamics,
2692: Europhysics Lett. 51, 1-7, 2000.
2693:
2694: \bibitem{Ogata4} Ogata, Y., Space-time point process models for earthquake
2695: occurrences, Ann. Inst. stat. Math. 50, 379-402, 1998.
2696:
2697: \bibitem{Grassproca} Grassberger, P. and Procaccia, I.,
2698: Phys. Rev. Lett. 50, 346-349, 1983.
2699:
2700: \bibitem{ScholzMandel} Scholz, C.H. and B.B. Mandelbrot,
2701: Fractals in geophysics, edited by Basel; Boston; Birkhauser Verlag, 1989.
2702:
2703: \bibitem{Anneexp} Sornette, A., P. Davy and D. Sornette,
2704: J. Geophys. Res. 98, 12111-12139, 1993.
2705:
2706: \bibitem{BartonLapointe1} Barton,~C.C. and La Pointe P.R., eds., Fractals in
2707: the Earth Sciences, Plenum Press, New York and London, 1995.
2708:
2709: \bibitem{Nanjo} Nanjo, K. and H. Nagahama, Spatial distribution of aftershocks
2710: and the fractal structure of active fault systems, Pure Appl. Geophys. 157, 575-588, 2000.
2711:
2712:
2713: \bibitem{Ouillonsor} Ouillon,~G. and Sornette,~D.,
2714: Unbiased multifractal analysis: application to fault patterns, Geophys. Res. Lett.
2715: 23,~3409-3412, 1996.
2716:
2717: \bibitem{Hamburger} Hamburger,~D., Biham,~O and Avnir,~D.,
2718: Apparent fractality emerging from models of random distributions,
2719: Phys. Rev. E 53, 3342-3358, 1996.
2720:
2721: \bibitem{Eneva} Eneva, M., Effect of limited data sets in evaluating the
2722: scaling properties of spatially distributed data -- An example from mining-induced
2723: seismic activity, Geophys. J. Int. 124, 773-786, 1996.
2724:
2725:
2726: \bibitem{Malcai} Malcai,~O., Lidar,~D.A., Biham,~O. and Avnir,~D.,
2727: Scaling range and cutoffs in empirical fractals, Phys. Rev. E 56,~2817-2828, 1997.
2728:
2729: \bibitem{HSG} Helmstetter, A, D. Sornette and J.-R. Grasso,
2730: Mainshocks are aftershocks of conditional foreshocks:
2731: How do foreshock statistical properties emerge from aftershock laws,
2732: in press in J. Geophys. Res.,
2733: (http://arXiv.org/abs/cond-mat/0205499) 2002.
2734:
2735: \bibitem{Montroll12} Montroll, E.W. and H. Scher, J. Stat. Phys., 9 (2),
2736: 101-135, 1973.
2737:
2738:
2739: \bibitem{Scher} Scher, H. and E.W. Montroll,
2740: Anomalous transit-time dispersion in amorphous solids, Phys.
2741: Rev. B 12, 2455, 1975.
2742:
2743: \bibitem{Kenkre} Kenkre,V.M., E.W. Montroll and M.F. Shlesinger,
2744: J. Stat. Phys. 9, 45, 1973.
2745:
2746: \bibitem{Shlesinger1} Shlesinger, M.F., Asymptotic solutions of
2747: continuous-time random walks, J. Stat. Phys. 10, 421-433, 1974.
2748:
2749: \bibitem{Weiss1} Weiss, G.H., Aspects and Applications of random
2750: Walk, North Holland, 1994.
2751:
2752: \bibitem{Hughes} Hughes, B.D. Random walks and Random Environments,
2753: Oxford University
2754: Press, 1995.
2755:
2756: \bibitem{Meltzner2} Metzler, R. and J. Klafter, The random walk's
2757: guide to anomalous
2758: diffusion:a fractional dynamics approach, Physics Reports 339,
2759: 1-77, 2000.
2760:
2761: \bibitem{Margolin} Margolin, G. and B. Berkowitz, Application of
2762: continuous time random walks to transport in porous media,
2763: J. Phys. Chem. 104, 3942-3947, 2000.
2764:
2765: \bibitem{Berko1} Berkowitz, B. and H. Scher, The role of
2766: probabilistic approaches to
2767: transport theory in heterogeneous media, Transport in Porous
2768: Media 42 (1-2), 241-263, 2001.
2769:
2770:
2771: \bibitem{Redner} Redner,~S., Random multiplicative processes: An elementary
2772: tutorial, Am. J. Phys., 58,~267-273, 1990.
2773:
2774: \bibitem{Schermarklaber} Scher, H. G. Margolin, R. Metzler, J. Klafter and
2775: B. Berkowitz, The dynamical foundation of fractal stream chemistry: the origin
2776: of extremely long retention times, accepted to Geophys. Res. Lett,
2777: e-print at cond-mat/0202326
2778:
2779: \bibitem{Kirchner} Kirchner, J.W., X. Feng and C. Neal, Fractal
2780: stream chemistry
2781: and its implications for contaminant transport in catchments, Nature
2782: 403, 524-527, 2000.
2783:
2784: \bibitem{barkai3} Barkai, E., R. Metzler and J. Klafter, From
2785: continuous time random walks to the fractional Fokker-Planck
2786: equation, Phys. Rev. E 61,
2787: 132-138, 2000.
2788:
2789: \bibitem{Mathai} Mathai, A.M. and R.K.Saxena, the H-function with applications
2790: in statistics and other disciplines,
2791: Wiley Eastern Limited, New Delhi, 1978.
2792:
2793: \bibitem{Roman} Roman, H.E. and P.A. Alemany, Continuous-time random
2794: walks and the fractional diffusion equation, J. Phys. A 27, 3407-3410, 1994.
2795:
2796: \bibitem{barkai1} Barkai, E., Fractional Fokker-Planck equation,
2797: solution and application,
2798: Phys. Rev. E 63, 046118, 2001a.
2799:
2800: \bibitem{Budde} Budde, C., D. Prato and M. Re,
2801: Superdiffusion in decoupled continuous time random walks,
2802: Phys. Lett. 283, 309-312, 2001.
2803:
2804: \bibitem{Kot1} Kotulski, M., Asymptotic distributions of
2805: continuous-time random walks -
2806: - A probabilistic approach, J. Stat. Phys. 81, 777-792, 1995.
2807:
2808: \bibitem{barkai2} Barkai, E., CTRW pathways to the fractional
2809: diffusion equation, preprint cond-mat/0108024, 2001b.
2810:
2811: \bibitem{Saichev} Saichev, A.I. and G.M. Zaslavsky, Fractional
2812: kinetic equations:
2813: solution and applications, Chaos 7 (4), 753-764, 1997.
2814:
2815: \bibitem{BeOr} Bender, C. M. and S. Orzag,
2816: Advanced Mathematical Methods for Scientists and Engineers,
2817: Mc. Graw-Hill, 1978.
2818:
2819: \bibitem{Weiss2} Weiss, G.H. and P.P. Calabrese, Occupation times of
2820: a CTRW on a lattice with anomalous sites, Physica A 234, 443-454, 1996.
2821:
2822: \bibitem{Sokolov1} Sokolov, I.M., A. Blumen and J. Klafter, Linear
2823: response in complex systems: CTRW and the fractional Fokker-Planck equations,
2824: preprint cond-mat/0107632, 2001.
2825:
2826: \bibitem{Sokolov2} Sokolov, I.M., Levy flights from a continuous-time process,
2827: Phys. Rev. E 63, 011104, 2001.
2828:
2829: \bibitem{Bender} Bender, C. M., S. Boettcher and P.N. Meisinger,
2830: Universality in random-walk
2831: models with birth and death, Phys. Rev. Lett. 75, 3210-3213, 1995.
2832:
2833:
2834: \bibitem{Shaw} Shaw B.E., Generalized Omori law for aftershocks and foreshocks
2835: from a simple dynamics, Geophys. Res. Lett. 20, 907-910, 1993.
2836:
2837:
2838: \bibitem{Marsan1} Marsan, D., C.J. Bean, S. Steacy and J. McCloskey,
2839: Observation of
2840: diffusion processes in earthquake populations and implications for the
2841: predictability of seismicity systems, J. Geophys. Res. 105,
2842: 28,081-28,094, 2000.
2843:
2844:
2845: \bibitem{Marsan2} Marsan, D., C.J. Bean, S. Steacy and J. McCloskey,
2846: Spatio-temporal
2847: analysis of stress diffusion in a mining-induced seismicity system.
2848: Geophys. Res. Lett. 26, 3697-3700, 1999.
2849:
2850:
2851: \bibitem{Marsan3} Marsan, D. and C.J. Bean, Average dynamical
2852: seismicity changes following
2853: a stress perturbation, for a world-wide catalogue,
2854: Submitted to Geophys. J. Int., 2001;
2855:
2856: \bibitem{Sotolongo} Sotolongo-Costa, O., J.C. Antoranz, A. Posadas, F. Vidal
2857: and A. V\'asquez, L\'evy flights and earthquakes, Geophys. Res. Lett.
2858: 27, 1965-1968, 2000.
2859:
2860: \end{thebibliography}
2861:
2862: \newpage
2863:
2864: \begin{table}[]
2865: \begin{center}
2866: \begin{tabular}{|c|c|c|}
2867: & ETAS & CTRW \\ \hline
2868: $\Psi(t)$ & pdf for a ``daughter'' to be born at time $t$ & pdf of
2869: waiting times \\
2870: & from the mother that was born at time $0$ & \\ \hline
2871: $\Phi(\vec r)$ & pdf for a daughter to be triggered & pdf of jump sizes \\
2872: & at a distance $\vec r$ from its mother & \\ \hline
2873: $m$ & earthquake magnitude & tag associated with each jump \\ \hline
2874: $\rho(m)$ & number of daughters & local branching ratio \\
2875: & per mother of magnitude $m$ & \\ \hline
2876: $n$ & average number of daughters created per mother & control
2877: parameter of the random \\
2878: & summed over all possible magnitudes & walk survival
2879: (branching ratio) \\ \hline
2880: $n < 1$ & subcritical aftershock regime & subcritical ``birth and
2881: death'' \\ \hline
2882: $n = 1$ & critical aftershock regime & the standard CTRW \\ \hline
2883: $n > 1$ & supercritical exponentially & explosive regime of the \\
2884: & growing regime & ``birth and death'' CTRW \\ \hline
2885: $N(t, \vec r)$ & number of events of any possible & pdf of just having \\
2886: & magnitude at $\vec r$ at time $t$ & arrived at $\vec r$ at time $t$ \\ \hline
2887: $W(t, \vec r)$ & pdf that an event at $\vec r$ has occurred at a
2888: time $t' \leq t$
2889: & pdf of being at $\vec r$ at time $t$ \\
2890: & and that no event occurred anywhere from $t'$ to $t$ &
2891: \end{tabular}
2892: \vspace{5mm}
2893: \caption{\label{table1} Correspondence between the ETAS
2894: (Epidemic-type aftershock sequence)
2895: and CTRW (continuous-time random walk) models. `pdf' stands for
2896: probability density
2897: function.
2898: }
2899: \end{center}
2900: \end{table}
2901: \newpage
2902:
2903: \begin{table}[]
2904: \begin{center}
2905: \begin{tabular}{|l|l|l|}
2906: & large $z$ & small $z$ \\
2907: & $r \ll D~t^H$ & $r \gg D~t^H$ \\ \hline
2908: $\mu<0.5$ & $p=1+\theta$ & $p=1-2\theta$\\
2909: $0.5 \leq \mu<2$ & $p=1-\theta+\theta/\mu$ & $p=1-2\theta$ \\
2910: $2 \leq \mu$ & $p=1-\theta/2$ & not defined \tablenotemark\\
2911: \end{tabular}
2912: \tablenotetext{The Omori exponent is not defined in this case because the
2913: dependence of $N(t,\vec r)$ with respect to time
2914: given by expression (\ref{exp1}) and represented in Figure \ref{nrtmus2} has
2915: a contribution from the exponential asymptotics which is different
2916: from a power-law for large distances $r \gg D~t^H$.}
2917: \vspace{5mm}
2918: \caption{\label{table2} Asymptotic values of the (renormalized) Omori
2919: exponent (of the dressed propagator)
2920: in the different regimes for $z\ll 1$ and $z\gg 1$ where $z \equiv {
2921: D~t^{H} \over r}$.
2922: }
2923:
2924: \end{center}
2925: \end{table}
2926: \newpage
2927:
2928:
2929: %FIGURE 1
2930: \begin{figure}
2931: \includegraphics[width=16cm]{Fig1map2811.2.eps} %jgrmap.2.dat N=30000 mu=1
2932: % programme ETAS2/mapjgr.m
2933: \caption{\label{map} Maps of seismicity generated by the ETAS model
2934: with parameters
2935: $b=1$, $\theta=0.2$, $\mu=1$, $d=1$ km, $\alpha=0.5$, $c=0.001$ day and a
2936: branching ratio $n=1$.
2937: The mainshock occurs at the origin of space with magnitude $M=7$. The minimum
2938: magnitude is fixed at $m_0=0$.
2939: The distances between mainshock and aftershocks follow a power-law
2940: with parameter
2941: $\mu=1$ and the local (or bare) Omori law is $\propto 1/t^{1+\theta}$.
2942: According to the theory developed
2943: in the text, the average distance between the first
2944: mainshock and the aftershocks is thus expected to grow as $R \sim t^{H}$
2945: with $H=0.2$ (equation (\ref{mgjgkr})).
2946: The two plots are for different time periods of the same numerical
2947: simulation, such that the same number of earthquakes $N=3000$ is obtained
2948: for each graph: (a) time between $0$ and $0.3$ days;
2949: (b) time between $30$ and $70$ yrs.
2950: Real aftershock sequences are indeed observed to last decades
2951: up to a century. Large black dots indicate large
2952: aftershocks around which other secondary aftershocks cluster.
2953: The mainshock is shown by a black star. At early times, aftershocks
2954: are localized close to the mainshock, and then diffuse and cluster around the
2955: largest aftershocks.
2956: }
2957: \end{figure}
2958:
2959: \clearpage
2960:
2961: %FIGURE 2
2962: \begin{figure}
2963: \includegraphics[width=16cm]{Fig5Nt-2.eps}
2964: \caption{\label{n} Seismicity rate $N(t)$ for the temporal ETAS model
2965: calculated for $\theta=0.3$ and $c=0.001$ day. The local law $\phi(t) \propto
2966: 1/t^{1+\theta}$, which gives the
2967: probability distribution of times between an event and its
2968: (first-generation) aftershocks is
2969: shown as a dashed line.
2970: The global law $N(t)$, which includes all secondary and successive
2971: aftershocks generated
2972: by all the aftershocks of the first event, is shown as a solid line for the
2973: three regimes, $n<1$, $n=1$ and $n>1$.
2974: In the critical regime $n=1$, the seismicity rate follows a
2975: renormalized or dressed
2976: Omori law $\propto 1/t^p$ for $t>c$ with an
2977: exponent $p=1-\theta$, smaller than the exponent of the local law $1+\theta$.
2978: In the sub-critical regime ($n<1$), there
2979: is a crossover from an Omori law $1/t^{1-\theta}$ for $t<t^*$
2980: to $1/t^{1+\theta}$ for $t>t^*$. In the super critical regime
2981: ($n>1$), there
2982: is a crossover from an Omori law $1/t^{1-\theta}$ for $t<t^*$ to an
2983: exponential increase $N(t) \sim \exp(t/t^*)$ for $t>t^*$. We have chosen
2984: on purpose values of $n=0.9997 <1$ and $n=1.0003 > 1$ very close to $1$
2985: such that the crossover time $t^* =10^9$ days given by (\ref{tstar}) is
2986: very large. In real data, such large $t^*$ would be undistinguishable from
2987: an infinite value corresponding to the critical regime $n=1$.
2988: This representation
2989: is chosen for pedagogical purpose to make clear the different
2990: regimes occurring
2991: at times smaller and larger than $t^*$. In reality, we can expect $n$ to
2992: be significantly smaller or larger than $1$, such that $t^*$ becomes
2993: maybe of the order of months, years to decades and the observed
2994: Omori law
2995: will thus lie in the cross-over regime, given an apparent Omori exponent
2996: anywhere from $1-\theta$ to $1+\theta$.
2997: }
2998: \end{figure}
2999: \clearpage
3000: %FIGURE 3
3001: \begin{figure}
3002: \includegraphics[width=16cm]{Fig2D1.eps}
3003: \caption{\label{cordim} Plot of the correlation function of the $3.000$
3004: epicenters generated in the time interval $[30, 70]$ yrs
3005: and shown in the right panel of figure \ref{map}, calculated
3006: following Grassberger-Procaccia's algorithm \cite{Grassproca},
3007: as a function of scale $r$, in double-logarithmic scales.}
3008: \end{figure}
3009:
3010: \clearpage
3011:
3012: %FIGURE 4
3013: \begin{figure}
3014: \includegraphics[width=9cm]{Fig6Ntrmus2t0.2.eps} % DIF/pcfdifmus2.m
3015: \caption{\label{nrtmus2} Rate of seismicity $N(t,r)$ in the critical
3016: regime $n=1$
3017: for $\theta=0.2$, $\mu>2$, $c'=1$ day and $\sigma=1$ km,
3018: evaluated from expressions (\ref{qdsi}) and (\ref{exp1}), plotted as a function
3019: of the time (a) for different values of the distance $r$ between the mainshock
3020: and its aftershocks, and (b,c) as a function of $r$ (logarithmic scale
3021: for $r$ in (b) and linear scale for $r$ in (c)) for different values of
3022: the time between the mainshock and its aftershocks.
3023: The temporal decay of seismicity with time is characterized by a
3024: power-law decay
3025: $ N(r,t) \sim {1 / t^{1-\theta/2}} $ close to the mainshock epicenter or at
3026: large times for $r \ll D t^{\theta/2}$.
3027: For large distances $r \gg D t^{\theta/2}$, there is a truncation of the
3028: power-law decay at early times $t^{\theta / 2} \ll r/D$, because the
3029: seismicity
3030: has not yet diffused up to the distance $r$.
3031: Although the distribution of distances between a mainshock and its
3032: direct aftershocks
3033: $\Phi(r)$ follows a power-law distribution with exponent $1+\mu$,
3034: the log-linear
3035: graph (c) shows that the global
3036: rate of aftershocks $N(\vec r,t)$ decreases approximately
3037: exponentially as a function of
3038: the distance from the mainshock, with a characteristic distance
3039: that increases with time.
3040: This is in agreement with expression (\ref{exp1}) which predicts
3041: $N(t, r) \sim \exp \left[\left( |\vec r| / D t^{\theta/2}
3042: \right)^{2 \over 2-\theta}\right]$,
3043: i.e., $N(t, r) \sim \exp \left( C(t) |\vec r|^q \right)$ with an
3044: exponent $q=2 /(2-\theta)$ close to $1$ within the exponential. The
3045: same remark as for
3046: figure \ref{n} applies: the representation of our predictions for
3047: very large times
3048: is made for pedagogical purpose to illustrate clearly the different regimes.
3049: }
3050: \end{figure}
3051:
3052:
3053:
3054: \clearpage
3055:
3056: %FIGURE 7
3057: \begin{figure}
3058: %\psfig{file=HSim811-1.1.ps ,width=16cm}
3059: %\psfig{file=Hdifmu2.5.eps ,width=16cm} %dif1302-1
3060: \includegraphics[width=16cm]{Fig7dif1203-1H.eps} %dif1203-1
3061: \caption{\label{poisq} Average distance between the first mainshock and its
3062: aftershocks as a function of the time from the mainshock, for numerical
3063: simulations of the ETAS model in the critical regime $n=1$, generated with the
3064: parameters $\theta=0.2$, $d=1$ km, $\mu=3$ and $c=10^{-3}$ day.
3065: The theoretical prediction for the diffusion exponent is thus $H=\theta/2=0.1$.
3066: We observe a crossover from a larger exponent at early times when the
3067: mean distance
3068: is close to the characteristic scale $d=1$ km of the distribution of
3069: distances between
3070: an aftershock and its progenitor,
3071: to a sub-diffusion with an exponent close to the theoretical prediction at
3072: large times. The solid line is a fit of the numerical data for times $t>10$ days,
3073: which gives an exponent $H=0.12$ slightly larger than the predicted
3074: value $H=0.1$.
3075: }
3076: \end{figure}
3077:
3078: \clearpage
3079:
3080: %FIGURE 8
3081: \begin{figure}
3082: %\psfig{file=Hdifmu0.5.eps ,width=16cm} % dif1302-2
3083: \includegraphics[width=16cm]{Fig8dif1103-1H.eps}
3084: \caption{\label{poimlksq} Average distance between the first mainshock and its
3085: aftershocks as a function of the time from the mainshock, for a numerical
3086: simulation of the ETAS model in the critical regime $n=1$, with
3087: $\theta=0.2$, $\mu=0.9$, $c'=1$ day and $d=1$ km.
3088: The solid line is a fit of the data which gives an exponent $H=0.25$ in good
3089: agreement with the predicted value $H=0.22$.}
3090: \end{figure}
3091:
3092: \clearpage
3093:
3094: %FIGURE 9
3095: \begin{figure}
3096: \includegraphics[width=12cm]{Fig9Ntrmu0.2t0.2.eps}
3097: \caption{\label{piuonk} Rate of seismicity $N(t,r)$ for $\theta=0.2$,
3098: $\mu=0.2$,
3099: $c'=1$ day and $\sigma=1$ km,
3100: evaluated from expressions (\ref{njhaagrssjb}) and (\ref{lqps1}),
3101: plotted as a function of the time (a) for different values of the distance
3102: $r$ between the mainshock and its aftershocks, and (b) as a function of $r$
3103: for different values of the time between the mainshock and its aftershocks.
3104: We stress again that the time scales shown here do not necessarily correspond
3105: to real observable time scales but are presented to demonstrate clearly the
3106: existence of the two regimes. The dashed lines give the predicted
3107: asymptotic dependence in each regime.
3108: }
3109: \end{figure}
3110:
3111: \clearpage
3112:
3113: %FIGURE 10
3114: \begin{figure}
3115: \includegraphics[width=12cm]{Fig10Ntrmu0.9t0.2.eps}
3116: \caption{\label{piuon} Rate of seismicity $N(t,r)$ for $\theta=0.2$, $\mu=0.9$,
3117: $c'=1$ day and $\sigma=1$ km, evaluated from expressions (\ref{njhaagrssjb}) and (\ref{lqps1}),
3118: plotted as a function of the time (a) for diffe
3119: rent values of the distance $r$
3120: between the mainshock and its aftershocks, and (b) as a function of
3121: $r$ for different
3122: values of the time between the mainshock and its aftershocks.
3123: The dashed lines give the predicted
3124: asymptotic dependence in each regime.
3125: }
3126: \end{figure}
3127:
3128: %FIGURE 11
3129: \begin{figure}
3130: %\psfig{file=Hdif1202-23.ps ,width=12cm}
3131: % Sim111-1.dat
3132: \includegraphics[width=10cm]{Fig11dif1103-1Ntr.eps}
3133: \caption{\label{poiq}
3134: Rate of seismicity $N(t,r)$ obtained from numerical simulations of
3135: the ETAS model generated
3136: with the same parameters as in Figure \ref{piuon} ($\theta=0.2$, $\mu=0.9$,
3137: $c'=1$ day and $d=1$ km).
3138: $N(r,t)$ is computed by averaging over 500 numerical
3139: realizations of the ETAS model. (a) aftershock rate as a function of the
3140: time from the mainshock for several distances
3141: $|\vec r|$ ranging from $0.01$ to $10^4$ km. (b)
3142: Apparent Omori exponent measured for times $t>10$
3143: as a function of the distance from the mainshock.
3144: The aftershock decay rate (with time) is larger close to the
3145: mainshock epicenter
3146: than at large distances from the mainshock. The asymptotic values for small and
3147: large distances are in agreement with the predictions (\ref{leqps5}) for
3148: $r \ll D t^{\theta/\mu}$ and (\ref{gjjsls}) for $r \gg D
3149: t^{\theta/\mu}$, which
3150: are shown as the horizontal dashed lines. (c) Rate of seismicity $N(t,r)$
3151: as a function of the distance between aftershocks and mainshock for various
3152: times. The theoretical prediction for large distances is shown as
3153: the dashed line
3154: with slope $-(1+\mu)$.
3155: }
3156: \end{figure}
3157:
3158: \clearpage
3159: \clearpage
3160: %FIGURE 12
3161: \begin{figure}
3162: \includegraphics[width=10cm]{Fig12H.eps}
3163: \caption{\label{H} Classification of the different regime of the diffusion
3164: of aftershocks in space as a function of time from the main shock.
3165: The bare Omori law for aftershocks decay with time as $1/t^{1+\theta}$.
3166: The jump size distribution between the earthquake
3167: ``mother'' and its ``daughters'' is proportional to
3168: $1/r^{1+\mu}$. $R(t)$ is the average distance between all
3169: aftershocks triggered up to time $t$ after the mainshock.}
3170: \end{figure}
3171:
3172: \end{document}
3173:
3174: