arxiv:physics/ v2 [physics.geo-ph] 18 Aug 2003

Similar documents
Impact of earthquake rupture extensions on parameter estimations of point-process models

A GLOBAL MODEL FOR AFTERSHOCK BEHAVIOUR

arxiv:physics/ v1 6 Aug 2006

The largest aftershock: How strong, how far away, how delayed?

arxiv:physics/ v1 [physics.geo-ph] 19 Jan 2005

Comparison of Short-Term and Time-Independent Earthquake Forecast Models for Southern California

Theory of earthquake recurrence times

Distribution of volcanic earthquake recurrence intervals

Aftershock From Wikipedia, the free encyclopedia

Interactions between earthquakes and volcano activity

Limitations of Earthquake Triggering Models*

Comment on Systematic survey of high-resolution b-value imaging along Californian faults: inference on asperities.

Comparison of short-term and long-term earthquake forecast models for southern California

Mechanical origin of aftershocks: Supplementary Information

Self-similar earthquake triggering, Båth s law, and foreshock/aftershock magnitudes: Simulations, theory, and results for southern California

Short-Term Properties of Earthquake Catalogs and Models of Earthquake Source

Because of its reputation of validity over a wide range of

Seismic Characteristics and Energy Release of Aftershock Sequences of Two Giant Sumatran Earthquakes of 2004 and 2005

Secondary Aftershocks and Their Importance for Aftershock Forecasting

The Role of Asperities in Aftershocks

Scaling of apparent stress from broadband radiated energy catalogue and seismic moment catalogue and its focal mechanism dependence

Journal of Asian Earth Sciences

On the Role of Multiple Interactions in Remote Aftershock Triggering: The Landers and the Hector Mine Case Studies

Probabilistic seismic hazard estimation in low-seismicity regions considering non-poissonian seismic occurrence

Observational analysis of correlations between aftershock productivities and regional conditions in the context of a damage rheology model

Southern California Earthquake Center Collaboratory for the Study of Earthquake Predictability (CSEP) Thomas H. Jordan

Does Aftershock Duration Scale With Mainshock Size?

Minimum preshock magnitude in critical regions of accelerating seismic crustal deformation

Aspects of risk assessment in power-law distributed natural hazards

Interpretation of the Omori law

A rate-state model for aftershocks triggered by dislocation on a rectangular fault: a review and new insights

Multifractal Analysis of Seismicity of Kutch Region (Gujarat)

Rotation of the Principal Stress Directions Due to Earthquake Faulting and Its Seismological Implications

UNIVERSITY OF CALGARY. Nontrivial Decay of Aftershock Density With Distance in Southern California. Javad Moradpour Taleshi A THESIS

High-resolution Time-independent Grid-based Forecast for M >= 5 Earthquakes in California

Small-world structure of earthquake network

Simulated and Observed Scaling in Earthquakes Kasey Schultz Physics 219B Final Project December 6, 2013

Accelerating energy release prior to large events in simulated earthquake cycles: implications for earthquake forecasting

Connecting near and farfield earthquake triggering to dynamic strain. Nicholas J. van der Elst. Emily E. Brodsky

Gutenberg-Richter Relationship: Magnitude vs. frequency of occurrence

Space-time ETAS models and an improved extension

Scale-free network of earthquakes

Stephen Hernandez, Emily E. Brodsky, and Nicholas J. van der Elst,

The Centenary of the Omori Formula for a Decay Law of Aftershock Activity

Strain rate and temperature dependence of Omori law scaling constants of AE data: Implications for earthquake foreshock-aftershock sequences

Mechanics of Earthquakes and Faulting

Strong foreshock signal preceding the L Aquila (Italy) earthquake (M w 6.3) of 6 April 2009

Adaptive Kernel Estimation and Continuous Probability Representation of Historical Earthquake Catalogs

ON NEAR-FIELD GROUND MOTIONS OF NORMAL AND REVERSE FAULTS FROM VIEWPOINT OF DYNAMIC RUPTURE MODEL

Magnitude Of Earthquakes Controls The Size Distribution Of Their. Triggered Events

Long lasting seismic swarm and pore pressure decrease

A mathematical formulation of accelerating moment release based on the stress accumulation model

Accelerating Seismic Crustal Deformation in the Southern Aegean Area

The Magnitude Distribution of Dynamically Triggered Earthquakes. Stephen Hernandez and Emily E. Brodsky

Modeling Aftershocks as a Stretched Exponential Relaxation

Space-time clustering of seismicity in California and the distance dependence of earthquake triggering

Scaling Relationship between the Number of Aftershocks and the Size of the Main

Chien-chih Chen Institute of Geophysics, National Central University, Chungli, Taiwan

Are aftershocks of large Californian earthquakes diffusing?

(1) Istituto Nazionale di Geofisica e Vulcanologia, Roma, Italy. Abstract

Quantifying the effect of declustering on probabilistic seismic hazard

Y. Y. Kagan and L. Knopoff Institute of Geophysics and Planetary Physics, University of California, Los Angeles, California 90024, USA

Quantifying early aftershock activity of the 2004 mid-niigata Prefecture earthquake (M w 6.6)

Toward interpretation of intermediate microseismic b-values

SUPPLEMENTARY INFORMATION

Assessing the dependency between the magnitudes of earthquakes and the magnitudes of their aftershocks

Finite data-size scaling of clustering in earthquake networks

Di#erences in Earthquake Source and Ground Motion Characteristics between Surface and Buried Crustal Earthquakes

Testing aftershock models on a time-scale of decades

ALM: An Asperity-based Likelihood Model for California

Earthquake predictability measurement: information score and error diagram

Interactions and triggering in a 3-D rate-and-state asperity model

Oceanic Transform Fault Seismicity Earthquakes of a Different Kind

Originally published as:

Long-Range Triggered Earthquakes That Continue After the Wavetrain Passes

THE DOUBLE BRANCHING MODEL FOR EARTHQUAKE FORECAST APPLIED TO THE JAPANESE SEISMICITY

Long-range triggered earthquakes that continue after the wave train passes

Power Law Distributions of Offspring and Generation Numbers in Branching Models of Earthquake Triggering

Accuracy of modern global earthquake catalogs

New Challenges In Earthquake Dynamics: Observing And Modelling A Multi-Scale System

PostScript file created: June 11, 2011; time 985 minutes RANDOM STRESS AND OMORI S LAW. Yan Y. Kagan

Potency-magnitude scaling relations for southern California earthquakes with 1.0 < M L < 7.0

Second Annual Meeting

Earthquake clusters in southern California I: Identification and stability

On the validity of time-predictable model for earthquake generation in north-east India

arxiv:physics/ v1 [physics.geo-ph] 4 Apr 2004

A TESTABLE FIVE-YEAR FORECAST OF MODERATE AND LARGE EARTHQUAKES. Yan Y. Kagan 1,David D. Jackson 1, and Yufang Rong 2

Seismic Quiescence before the 1999 Chi-Chi, Taiwan, M w 7.6 Earthquake

Aging and scaling of aftershocks

Earthquakes Chapter 19

Introduction The major accomplishment of this project is the development of a new method to identify earthquake sequences. This method differs from

Supporting Information for Break of slope in earthquake-size distribution reveals creep rate along the San Andreas fault system

Recurrence Times for Parkfield Earthquakes: Actual and Simulated. Paul B. Rundle, Donald L. Turcotte, John B. Rundle, and Gleb Yakovlev

Afterslip, slow earthquakes and aftershocks: Modeling using the rate & state friction law

Space time ETAS models and an improved extension

Delayed triggering of microearthquakes by multiple surface waves circling the Earth

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 21 Jan 2004

Anomalous early aftershock decay rate of the 2004 Mw6.0 Parkfield, California, earthquake

RAPID SOURCE PARAMETER DETERMINATION AND EARTHQUAKE SOURCE PROCESS IN INDONESIA REGION

A global classification and characterization of earthquake clusters

Transcription:

Is Earthquake Triggering Driven by Small Earthquakes? arxiv:physics/0210056v2 [physics.geo-ph] 18 Aug 2003 Agnès Helmstetter Laboratoire de Géophysique Interne et Tectonophysique, Observatoire de Grenoble, Université Joseph Fourier, France (Dated: October 27, 2018) Using a catalog of seismicity for Southern California, we measure how the number of triggered earthquakes increases with the earthquake magnitude. The trade-off between this relation and the distribution of earthquake magnitudes controls the relative role of small compared to large earthquakes. We show that seismicity triggering is driven by the smallest earthquakes, which trigger fewer events than larger earthquakes, but which are much more numerous. We propose that the nontrivial scaling of the number of triggered earthquakes emerges from the fractal spatial distribution of seismicity. Large shallow earthquakes are always followed by aftershocks, that are due the stress change of the mainshock. The number n(m) of aftershocks of a mainshock of magnitude M has been proposed to scale with M as [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14] n(m) 10 αm. (1) This relation accounts for the fact that large earthquakes trigger many more aftershocks than small earthquakes. A similar relation holds for the distribution of earthquake magnitudes P(M) [15] given by P(M) 10 bm, (2) withbtypicallycloseto1, whichimplies thatsmallearthquakes are much more frequent than large earthquakes. Because large earthquakes release more energy and trigger more aftershocks than smaller earthquakes, it is usually accepted that interactions between earthquakes and earthquake triggering are dominated by the largest earthquakes. However, because they are much more frequent that larger earthquakes, small earthquakes are also just as important as large earthquakes in redistributing the tectonic forces if b = 1 [16]. Other quantities, such as the Benioff strain ǫ.75m, are dominated by small earthquakes if b > 0.75. The α-exponent is an important parameter of earthquake interaction that is used in many stochastic models of seismicity or prediction algorithms [2, 3, 4, 5, 7, 8, 14, 17]. This parameter controls the relative role of small compared to large earthquakes. While there is a significant amount of literature on the b-value, very few studies have measured accurately the α exponent in real seismicity data. Many studies use α = b without justification [2, 8, 10, 14, 17]. In this case, small earthquakes are just as important as larger ones for the triggering process. Using (1) and (2), the global number N(M) of aftershocks triggered by all earthquakes of magnitude M scales as N(M) = P(M) n(m) 10 (α b)m, (3) and is indeed independent of M in the case α = b. In the case α < b, aftershock triggering is controlled by the smallest earthquakes, while the largest earthquakes dominate if α > b. A few studies measured directly α from aftershocks sequences, using a fit of the total number of aftershocks as a function of the mainshock magnitude [6, 9, 11, 13]. These studies yield α-value close to 1, but the limited range of the mainshock magnitude considered and the large scatter of the number of aftershocks per mainshock do not allow an accurate estimation of α. The case α = b also explains another well documented property of aftershocks, known as Bath s Law [13, 14, 18], which states that the difference between the mainshock magnitude and its largest aftershock is on average close to 1.2, independently of the mainshock magnitude. Again, the limited range of mainshock magnitudes used in these studies and possible biases of data selection [19] does not allow to test the dependence of the magnitude difference as a function of the mainshock magnitude. Other studies measured α indirectly using a stochastic triggering model called Epidemic Type Aftershock Sequence model (ETAS) [3, 7, 20] based only on the Gutenberg-Richter and Omori laws [2, 7]. This model assumes that each earthquake above a magnitude threshold m 0 can trigger direct aftershocks, with a rate that increases as 10 αm with its magnitude, and decays with time according to Omori law[21]. The average total number of aftershocks n(m), including the cascades of indirect aftershocks, has the same dependence 10 αm with the mainshock magnitude M as the number of direct aftershocks. Using this model, α can be measured using a maximum likelihood method [3, 7, 20]. For instance, [20] analyzed 34 aftershock sequences in Japan and measured α in the range [0.2 1.9] with a mean value of 0.86. The α-values obtained from the inversion of this model are not well constrained due to the small number of events available and to possible biases of the inversion method. Indeed, these studies do not take into account the influence of earthquakes below the detection threshold, which may significantly bias the estimation of α. This method may also be biased by the incompleteness of the catalog just after the mainshock, and by possible trade-offs between the ETAS parameters. The regime α b of the

2 ETAS model is probably not relevant for real seismicity. If we do not assume a roll-off of the magnitude distribution P(M) for large M, this regime gives a finite time singularity of the seismicity rate which goes to infinity in finite time t c as 1/(t c t) m [23]. Such a power-law increase of seismic activity can describe the acceleration of the deformation preceding material failure as well as a starquake sequence [23], but cannot describe a stationary seismic activity. In this study, we use a stacking method to estimate the average rate of earthquakes triggered (directly or indirectly) by a previous earthquake as a function of the magnitude of the triggering earthquake. We use the seismicity catalog of Southern California provided by the Southern California Data Center [24], which covers the time period 1975-2003, and which is complete above M = 3 for this time period. The magnitude distribution shown in Fig. 1b follows the Gutenberg-Richter law for M 3 and attests for the completeness of the catalog for this time period above magnitude 3. We do not use the usual distinction between foreshocks, mainshocks and aftershocks, and the constraint that the aftershocks must be smaller than the mainshock because this classification is not based on physical differences. Indeed, recent studies have shown that a simple model that assumes that each earthquake can trigger earthquakes of any magnitude, without any distinction between foreshocks, mainshocks and aftershocks can reproduce many properties of real seismicity including realistic foreshock sequences[2, 3, 7, 14, 22, 25]. Constraining triggered earthquakes to be smaller than the mainshock would obviously underestimate the number of earthquakes triggered by small mainshocks and thus overestimate α. We define a triggered earthquake as any event occurring in a space-time window R T after a preceding mainshock above the background level, whatever the relative magnitude of the triggered and triggering earthquakes. We consider as a potential mainshock each earthquake that has not been preceded by a previous larger earthquake in a space-time window d T in order to estimate the rate of seismicity triggered by this mainshock removing the influence of previous earthquakes. This definition of triggered earthquakes and mainshocks contains unavoidably a degree of arbitrariness in the choice of the space-time windows but the estimation of α is found to be robust when changing T, R and d. We have tested different methods for the choice of R, either fixed or increasing with the mainshock magnitude. We use a distance R increasing with the mainshock magnitude because the aftershock zone is usually found to scale with the rupture length [26, 27]. We use R equal to 1 rupture length of the mainshock. For small mainshock magnitudes, this choice would lead to unacceptable values of R smaller than the location error, and thus to underestimate the number of triggered earthquakes of small mainshocks. Therefore, weimpose R > 5km, largerthan the location error. Taking R fixed has the advantage of not introducing by hand any scaling between the aftershock zone and the mainshock magnitude. However, it may overestimate the number of earthquakes triggered by the smallest mainshocks if R is too large, or underestimate the number of triggered events of the largest mainshock if R is too small. The results obtained for T = 1 year, R = 0.01.5M km and d = 50 km are presented in Fig. 1. The rate of triggered earthquakes is found to decay according to Omori s law K(M)/t p, with the same exponent p 0.9 for all mainshock magnitudes M (Fig. 1a). The amplitude K(M)increasesexponentially 10 αm asafunction of M with α = 0.81 (Fig. 1b). This confirms that the scaling of the rate of triggered earthquakes with M follows (1). Our method is more accurate than previous studies [6, 9, 11, 13]. Indeed, these studies [6, 9, 11, 13] determine the scaling of n(m) with M using the total number of aftershocks [6, 9, 11, 13] in a time window [0 T] after a mainshock, and can thus be biased by the incompleteness of the catalogs just after a large mainshock or by the background seismicity at large times after a mainshock. In contrast, in order to deal with these problems, we use the seismicity rate in the time window where we observe the Omori law decay characteristic of triggered seismicity. The value of α is robust when increasing or decreasing the distance R used for the selection of triggered earthquakes between 1 to 5 rupture lengths, or when increasing the minimum value of R from 2 to 10 km. Selecting earthquakes within a disk of fixed radius R = 50 km for all mainshock magnitudes yields a slightly smaller value α = 0.72. Decreasing R leads to a smaller value of α because it underestimates the number of events triggered by the largest mainshocks, which have a rupture size larger than R. When increasing R from 10 to 100 km, the value of α first increases with R and then saturates around α = 0.72 for R 50 km. We have also checked that α is not sensitive to the parameter d used for the selection of mainshocks if d 50 km. All values of α, for reasonable values of the parameters R in the range 30 100 km, T between 0.1 and 2 yrs and d > 50 km, and for different time periods of the catalog, are in the range [0.7 0.9]. We have also tested the method on synthetic catalogs generated with the ETAS model. We recover the α parameter with an error smaller than 0.05. For the same catalog of seismicity, we measure using a maximum likelihood method the b-value of the Gutenberg-Richter law (2) equal to b = 1.08±0.10. We have also tested that the magnitude distribution P(M) of triggered events is independent of the mainshock magnitude. Fig. 2 shows the magnitude distribution of triggered events for different ranges of the mainshock magnitude, using the same data as in Fig. 1. This figure shows that a large earthquake can be triggered by a smaller

3 earthquake. Our results suggest that α is significantly smaller than the b exponent of the magnitude distribution. Whether the exponent α varies with region and maybe even with time is an interesting question that is outside the scope of this Letter but we urge further studies in that direction. We now propose a simple explanation for this nontrivial scaling of the number of triggered earthquakes with the mainshock magnitude, and we suggest that α can be related to the fractal structure of the spatial distribution of seismicity. It is widely accepted that the aftershock zone scales with the rupture length [26, 27]. While the area affected by the stress variation induced by an earthquake increases with the rupture length, the stress drop is independent of the mainshock magnitude [28, 29]. The stressvariationatadistancefromthe mainshock proportional to the fault length L is thus independent of the mainshock magnitude, neglecting the effect of the finite width of the crust and the visco-elastic deformation in the lower-crust. Therefore, assuming that earthquakes triggered by the stress change induced by the mainshock, the density of earthquakes triggered at a distance up to R L from the mainshock is independent of the mainshock magnitude. The increase of the number of triggered events with the mainshock magnitude results only from the increase in the aftershock zone size with the rupture length. The rupture length is usually related to the magnitude by [28] L.5M. (4) The same relation thus holds between the aftershock zone size R and the mainshock magnitude. In order to estimate the scaling of the number of triggeredevents with the rupture length, we need to make an assumption about the spatial distribution of earthquakes around the mainshock. Assuming that triggered earthquakes are uniformly distributed on the fault plane, and using(4), the number of earthquakes triggered by a mainshock of magnitude M is given by n(m) L 2 10 M [9] and thus leads to α = 1. The value α = 0.5 obtained for a numerical model of seismicity [12] suggests that in this model earthquakes are triggered mostly on the edge of the fracture area of the mainshock [12]. Our result α = 0.8 for Southern California seismicity implies that triggered earthquakes are distributed neither uniformly on the rupture plane nor on the edge of the rupture, but rather on a fractal structure of dimension D < 2. Using the definition of the capacity fractal dimension, the number of aftershocks is n(m) R D, (5) where R is the characteristic length of the aftershock zone. Using(4)and(5), weobtainthescalingofthenumber of triggered earthquakes with the mainshock magnitude n(m).5dm (6) which givesα = 0.5D. Ourestimation α = 0.8forSouthern California seismicity thus suggests D = 1.6. This value of the fractal dimension of aftershocks hypocenters has never been measured for Southern California seismicity. Our estimate of D is significantly smaller than the value measured in the range [2 2.8] for aftershock sequences in Japan [20]. This fractal dimension of the spatial distribution of triggered earthquakes results in part from the fractal structure of the fault system [30], but it may also reflect the non-uniformity of the distribution of the earthquakes on the fault due to the heterogeneity of stress or strength on the fault. The fractal dimension of the aftershock distribution may thus be smaller than the fractal distribution of the fault system. While the energy release and the total slip on faults is controlled by the largest earthquakes, the suggestion that α < b implies that small earthquakes may be more important than large earthquakes in triggering earthquakes. We have also checked that the magnitude distribution of triggered earthquakes is independent of the mainshock magnitude (Fig. 2). This implies that earthquake triggering is driven by the smallest earthquakes at all scales, even for the largest earthquakes. Other observations [25] support the conclusion that the same mechanisms can explain the triggering of a large earthquake by a smaller one and the triggering of a small earthquake by a previous larger event. Recent studies [14] have proposed that secondary aftershocks dominate an aftershock sequence, so that subsequent large aftershocks are more likely to be triggered indirectly by a previous aftershock of the mainshock. Our study further suggests that the smallest earthquakes will dominate the triggering of following earthquakes. The importance of small earthquakes casts doubts on the relevance of calculations of direct stress transfer functions to predict seismicity [31], because large earthquakes are likely to be triggered by the smallest earthquakes below the detection threshold of the seismic network. Small earthquakes taken individually have a very low probability of triggering a large earthquake. But because they are much more numerous than larger earthquakes, collectively, they trigger more earthquakes. This result requires the existence of a small magnitude cut-off m 0, below which earthquakes may occur but cannot trigger earthquakes larger than m 0, or a change of the scaling of N(M) given by (3) for small earthquakes, otherwise the seismicity at all scales would be controlled by infinitely small earthquakes. I am very grateful to D. Sornette, J.-R. Grasso, A. McGarr, L. Margerin, C. Voisin and T. Gilbert for useful suggestions and discussions. I am also grateful for the earthquake catalog data made available by the Southern California Earthquake Data Center.

4 Now at Institute of Geophysics and Planetary Physics, University of California Los Angeles [1] Utsu, T., J. Fac. Sci. Hokkaido Univ., Ser.VII 3, 129 (1969). [2] Kagan, Y. Y. and L. Knopoff, Science 236, 1563, (1987). [3] Kagan, Y. Y., Geophys. J. Int. 106, 135 (1991). [4] Reasenberg, P., J. Geophys. Res. 90 (1985). [5] Reasenberg, P. A., J. Geophys. Res. 104, 4755 (1999). [6] Singh, S. K. and G. Suarez, Bull. Seism. Soc. Am. 78, 230 (1988). [7] Ogata, Y., J. Am. stat. Assoc. 83, 9 (1988). [8] Reasenberg, P. A. and L. M. Jones, Science 243, 1173 (1989). [9] Yamanaka, Y. and K. Shimazaki, J. Phys. Earth 38 (1990). [10] Davis, S. D. and C. Frohlich, Geophys. J. Int. 104, 289 (1991). [11] Molchan, G. M. and O. E. Dmitrieva, Geophys. J. Int. 109, 501 (1992). [12] Hainzl, S., G. Zoller and J. Kurths, Nonlinear Processes in Geophysics 7, 21 (2000). [13] Drakatos, G. and J. Latoussakis, Journal of Seismology 5, 137 (2001). [14] Felzer, K. R. et al., J. Geophys. Res., 10.1029/2001JB000911 (2002). [15] Gutenberg, B. and C. F. Richter, Seismicity of the Earth and Associated Phenomena, Princeton Univ. Press, Princeton, NJ (1949). [16] Hanks, T. C., Science 256, 1430 (1992). [17] Console, R. and M. Murru, J. Geophys. Res. 166, 8699 (2001). [18] Bath, M., Tectonophysics 2, 483 (1965). [19] Vere-Jones, D. Bull. Seism. Soc. Am. 59, 1535 (1969). [20] Guo, Z. andy. Ogata, J. Geophys. Res. 102, 2857 (1997). [21] Omori, F., J. Coll. Sci. Imp. Univ. Tokyo 7, 111 (1894). [22] Helmstetter, A., D. Sornette and J.-R. Grasso, J. Geophys. Res. 108, 10.1029/2002JB001991 (2003). [23] Sornette, D. and A. Helmstetter, Phys. Rev. Lett. 89, 158501 (2002). [24] The catalog can be downloaded from www.scecdc.scec.org/ftp/catalogs/scsn. [25] Helmstetter, A. and D. Sornette, Foreshocks explained by cascades of triggered seismicity, in press in J. Geophys. Res. (2003) (physics/0210130). [26] Utsu, T., Geophys. Magazine 30, 521 (1961). [27] Kagan, Y. Y., Bull. Seism. Soc. Am. 92, 641, (2002). [28] Kanamori, H. and D. Anderson, Bull. Seism. Soc. Am. 65, 1073 (1975). [29] Ide, S. and G. C. Beroza, Geophys. Res. Lett. 28, 3349 (2001). [30] Bonnet, E. et al., Reviews of Geophysics 39, 347 (2001). [31] Stein, R.S., Nature, 402, 605 (1999).

5 seismicity rate (day 1 ) (a) 10 2 10 3 10 4 10 4 10 3 10 2 time after mainshock (days) rate of triggered earthquakes (o) 10 4 (b) 10 4 3 3.5 4 4.5 5 5.5 6 6.5 7 7.5 10 1 mainshock magnitude cumulative magnitude distribution (x) FIG. 1: Average rate of triggered earthquakes n(m,t) as a function of time t after the triggering earthquake (a) for different values of the magnitude M of the triggering earthquake increasing from 3 to 7 with a step of 0.5 from bottom to top. The rate of aftershocks K(M) as a function of M is shown in panel (b) (circles) with the cumulative magnitude distribution (crosses). K(M) is obtained by fitting each curve n(m, t) by K(M)/t 0.9 in the range 0.01 < t < 365 days for M < 6.5. For large M 6.5 mainshocks, there is a roll-off of the seismicity rate for small times after the mainshock due to the incompleteness of the catalog after large mainshocks, caused by the saturation of the seismic network. Therefore we measure K(M) in the range t > 0.1 day for M = 6.5 and t > 0.3 day for M = 7. cumulative magnitude distribution 10 2 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 10 3 3 3.5 4 4.5 5 5.5 6 6.5 7 magnitude FIG. 2: Cumulative magnitude distribution of triggered earthquakes for different values of the mainshock magnitude between 3 (dark line, small circles) and 7 (gray line, large symbols) using the same time interval for the selection of aftershocks as in Fig. 1.