Mechanical origin of aftershocks: Supplementary Information

Similar documents
Distribution of volcanic earthquake recurrence intervals

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

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

Theory of earthquake recurrence times

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

arxiv:physics/ v1 6 Aug 2006

A GLOBAL MODEL FOR AFTERSHOCK BEHAVIOUR

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

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

Journal of Asian Earth Sciences

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

Minimum preshock magnitude in critical regions of accelerating seismic crustal deformation

The Role of Asperities in Aftershocks

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

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

Limitations of Earthquake Triggering Models*

Finite data-size scaling of clustering in earthquake networks

Small-world structure of earthquake network

Scale-free network of earthquakes

Modeling Aftershocks as a Stretched Exponential Relaxation

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

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

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

Proximity to Past Earthquakes as a Least-Astonishing Hypothesis for Forecasting Locations of Future Earthquakes

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

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

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

Does Aftershock Duration Scale With Mainshock Size?

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

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

Aftershock From Wikipedia, the free encyclopedia

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

Triggering of Aftershocks of the Japan 2011 Earthquake by Earth Tides

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

Earthquake clusters in southern California I: Identification and stability

Because of its reputation of validity over a wide range of

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

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

RELOCATION OF THE MACHAZE AND LACERDA EARTHQUAKES IN MOZAMBIQUE AND THE RUPTURE PROCESS OF THE 2006 Mw7.0 MACHAZE EARTHQUAKE

Performance of national scale smoothed seismicity estimates of earthquake activity rates. Abstract

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

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

Delayed triggering of microearthquakes by multiple surface waves circling the Earth

Geophysical Journal International

Application of a long-range forecasting model to earthquakes in the Japan mainland testing region

Edinburgh Research Explorer

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

Multifractal Analysis of Seismicity of Kutch Region (Gujarat)

Testing aftershock models on a time-scale of decades

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

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

Interactions between earthquakes and volcano activity

Are aftershocks of large Californian earthquakes diffusing?

log (N) 2.9<M< <M< <M< <M<4.9 tot in bin [N] = Mid Point M log (N) =

Identifying the occurrence time of an impending mainshock: a very recent case

Space-time ETAS models and an improved extension

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

29th Monitoring Research Review: Ground-Based Nuclear Explosion Monitoring Technologies

Codal provisions of seismic hazard in Northeast India

Quantifying the effect of declustering on probabilistic seismic hazard

SUPPLEMENTARY INFORMATION

arxiv: v1 [physics.geo-ph] 6 Jun 2016

The 150th Anniversary of Fusakichi Omori

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

High-precision location of North Korea s 2009 nuclear test

Adaptive Kernel Estimation and Continuous Probability Representation of Historical Earthquake Catalogs

Negative repeating doublets in an aftershock sequence

Aftershocks are well aligned with the background stress field, contradicting the hypothesis of highly heterogeneous crustal stress

PostScript file created: April 24, 2012; time 860 minutes WHOLE EARTH HIGH-RESOLUTION EARTHQUAKE FORECASTS. Yan Y. Kagan and David D.

Interpretation of the Omori law

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

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

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

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

arxiv:cond-mat/ v1 17 Jan 2003

Aging and scaling of aftershocks

Are Declustered Earthquake Catalogs Poisson?

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

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

Geophysical Journal International

Statistical Properties of Marsan-Lengliné Estimates of Triggering Functions for Space-time Marked Point Processes

A forward test of the Decelerating-Accelerating Seismic Strain model in the Mediterranean

Onto what planes should Coulomb stress perturbations be resolved?

Exploring aftershock properties with depth

Toward automatic aftershock forecasting in Japan

Estimation of S-wave scattering coefficient in the mantle from envelope characteristics before and after the ScS arrival

Coulomb stress changes due to Queensland earthquakes and the implications for seismic risk assessment

Modeling of magnitude distributions by the generalized truncated exponential distribution

RELOCATION OF LARGE EARTHQUAKES ALONG THE PHILIPPINE FAULT ZONE AND THEIR FAULT PLANES

The above-average rate of earthquakes of magnitude 8 and

COULOMB STRESS CHANGES DUE TO RECENT ACEH EARTHQUAKES

JCR (2 ), JGR- (1 ) (4 ) 11, EPSL GRL BSSA

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

Comment on the paper. Layered seismogenic source model and probabilistic seismic-hazard analyses in. Central Italy

Earthquake Clustering and Declustering

Short-Term Earthquake Forecasting Using Early Aftershock Statistics

Earthquake Patterns in Diverse Tectonic Zones of the Globe

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

Seismic Hazard Assessment for Specified Area

Multi-dimensional residual analysis of point process models for earthquake. occurrences. Frederic Paik Schoenberg

Transcription:

Mechanical origin of aftershocks: Supplementary Information E. Lippiello Department of Mathematics and Physics, Second University of Naples, Via Vivaldi 43, 81100 Caserta, Italy & Kavli Institute for Theoretical Physics, Kohn Hall, University of California, Santa Barbara, CA 93106-4030 USA F. Giacco & C. Godano Department of Mathematics and Physics, Second University of Naples, Via Vivaldi 43, 81100 Caserta, Italy W. Marzocchi INGV, Via di Vigna Murata 605, 00143 Roma, Italy. L. de Arcangelis Department of Industrial & Information Engineering and CNISM, Second University of Naples, Aversa (CE), Italy & Kavli Institute for Theoretical Physics, Kohn Hall, University of California, Santa Barbara, CA 93106-4030 USA

2 TABLE I. Details on the experimental catalogs. Geographic area Temporal interval Latitude interval Longitude Interval Number of T min, T max Lat min, Lat max Lon min, Lon max m M > 6.5 mainshocks Southern California [1] 1/1/1984, 12/31/2002 31, 37.3-121.7, - 113.8 3 Northern California [2] 1/1/1984, 12/31/2011 35.3, 42.4 34-126, 1116 3 Japan mainland [3] 1/1/1966, 01/30/2011 30, 45.5 128.3, 147 6 Alaska [4] 1/1/1990, 01/31/2013 50, 71 130, 188 1 Italy [5] 1/1/2002, 12/31/2012 35, 48 6, 19 2 TABLE II. Parameters in the declustering procedure Measured quantity t 1 t 2 R c form M [2.5,4.5] 100 sec 1 day 0.02 km L a for m M [2.5,4.5] 0 sec 600 sec 10 0.5mM 2 km c,l a for m M 5.9 30 min 100 days 10 0.5mM 2 km I. DETAILS ON USED DATA SETS. The study on intermediate mainshock magnitudem M [2.5,4.5] is performed on the relocated catalog for Southern California [1]. Details on the catalogs used for the study on other geographic areas for large mainshocksm M > 6.5 can be found in Table 1. In Table 1 we also report the number of mainshocks identified for each data set. For the Italian region we select the two largest m M 5.9 mainshocks in the catalog. In all cases we restrict to events with depth smaller than 80 km. In Table 2 we report the parameters t 1,t 2 and R implemented in the aftershock identification procedure. We have used three different parameter sets, one for aftershock identification after large mainshocks, one for the measure of c in mainshock sequences with m m [2.5,4.5] and one for the measure of L a in mainshock sequences with m m [2.5,4.5]. The different choice of parameter sets is motivated by the observation that aftershock rate is much larger than the background seismicity for several days after large shocks. In particular, for m m > 6.5 the contamination by background events is very weak and one can consider also large temporal intervals extending up to t 2 = 100 days where the Omori law is clearly observed. Conversely, also in a temporal interval of few hours after small mainshocks m m 4.5, the contribution of background seismicity cannot be neglected. To fix this problem, since in the evaluation of c we are interested in the aftershock temporal organization, we used a sufficiently large temporal interval t 2 = 1 day but a narrow spatial range (R = 0.02 km). This is exactly the same spatiotemporal range considered in ref.[4]. On the other hand, in the evaluation ofl a we are interested in the aftershock distribution in space and in this case we have considered a larger R (R = l(m m )) but a very small temporal interval, t 2 = 10 minutes. Within these spatio-temporal ranges, we are quite confident that the influence of background events is never relevant. The choice of t 1 is motivated by data incompleteness in the first part of main-aftershock sequences (see Sec.III). The observed incompleteness is present on larger time scales for larger mainshocks and this implies different values oft 1 for different data sets.

3 II. THE ROLE OF THEλ-DEPENDENCE OF THEb-VALUE ON THEλ-DEPENDENCE OF THE AFTERSHOCK AREA This section is devoted to show that the observed dependence ofl a onλis not an indirect consequence of the dependence of the b-value on λ. Let us introduce the main-aftershock epicentral distance δr and and let us indicate with δr(m, λ) the average main-aftershock epicentral distance for a mainshock with magnitude m and rake angle λ. Let us consider two mainshocks m 1,m 2 with same magnitude m 1 = m 2 = m and different rake angles λ 1 λ 2. Under the assumption that the aftershock spatial distribution is λ independent, the dependence of δr(m,λ) on the rake angle is only through the mainshock magnitude. Therefore, sincem 1 = m 2, we would get δr(m 1,λ 1 ) = δr(m 2,λ 2 ) = Γ(m), whereγ(m) l(m) 10 0.5m is a monotonic increasing function of m [6]. By averaging now over mainshock magnitudes in the range m [2.5,4.5) we obtain δr(λ) = 4.5 2.5 dm δr(m, λ) P(m, λ) (1) where the mainshock magnitude distribution P(m) follows the Gutenberg-Richter law P(m) 10 b(λ)m. Under the above hypothesis of an aftershock spatial distribution independent of λ, we obtain δr(λ) 4.5 2.5 dmγ(m)10 b(λ)m. (2) Therefore the only dependence onλis throughb(λ) and, in this case, δr(λ) is a decreasing function ofb. Comparing the above prediction with data from real seismic catalogs (red open squares in Suppl.Fig.1) we find the opposite dependence: δr(λ) is an increasing function ofb. Experimental data therefore indicate that also δr(m,λ) explicitly depends onλand therefore also the aftershock spatial organization is affected by the rake angle. In order to separate the dependence on λ of δr(m,λ) from the one of b(λ), in the manuscript we have considered the quantity L a = 4.5 2.5 dm δr(m,λ) P(m,λ). (3) l(m) Indeed, from the experimental scaling relation Γ(m) l(m), it is reasonable to expect that δr(m, λ) /l(m) weakly depends on m and, since dmp(m,λ) = 1, L a is substantiallyb(λ) independent. The above arguments can be explicitly tested by means of numerical simulations. In particular we consider synthetic catalogs generated according to the Epidemic Type Aftershock Sequence (ETAS) model [7, 8]. This model describes seismic occurrence as the superposition of independent mainshocks (the ancestors) which trigger their offspring (the aftershocks). Furthermore, every aftershock can trigger its own offspring leading to an epidemic structure. Well established empirical laws are implemented for the main-aftershock spatio-temporal correlations. In particular here we use typical model parameters and spatial kernels as in ref. [9]. We assume that the only parameter depending onλisband all the other parameters areλindependent. We therefore generate different catalogs, each one corresponding to a different value of b. To each catalog we apply the main-aftershock selection procedure used for the experimental catalog. Results plotted in Suppl.Fig.1 (black circles in the upper panel) confirm our prediction that < δr > is a decreasing function of b, an opposite trend with respect to experimental results (red squares). In the lower panel of Suppl.Fig.1 we present the comparison between the dependence of L a on b in the numerical catalog (black circles) and the experimental one (red squares). As expected the numerical L a is substantially b independent indicating that the behavior of L a in experimental data is the correct measure of theλdependence of the aftershock area.

4 0,5 <δr> 0,4 0,3 0,2 0,1 L a 0,7 0,6 0,5 0,4 0,3 0,2 0,1 0,8 1 1,2 1,4 b value FIG. 1. (Upper Panel) Parametric plot of δr(λ) defined in Eq.(2) as function of theb-value for experimental results (red squares) and numerical data (black circles) from the ETAS model simulation. (Lower panel) The same as in the upper panel for the aftershocks area L a defined in Eq.(3) and discussed in the manuscript. III. THE EFFECT OF SHORT TERM AFTERSHOCK INCOMPLETENESS ON EXPERIMENTAL FINDINGS This section is devoted to show that the observed correlations amongbandcandl a, observed in experimental data sets, cannot be spurious effects related to catalog incompleteness. Because of technical problems experimental catalogs are incomplete in particular in the early part of aftershock sequences following large shocks [10]. This short term aftershock incompleteness (STAI) leads to deviations from the GR law. More precisely, one can define a crossover magnitude m c ( t), where t is the time from the mainshock so that P(m) follows the GR relation for m > m c ( t), whereas P(m) exhibits a flat behavior for m < m c ( t). The crossover magnitude m c ( t) exhibits a regular trend consistent with m c ( t) = m 0 δlog( t) m 1 (4) where m 0 is the mainshock magnitude, δ and m 1 are fitting parameters. Helmstetter et al. [11] find δ = 0.75 and m 1 = 4.5. Some authors interpret the c value for the onset of the power law decay in the Omori law as an artifact related to STAI. More precisely, indicating with m a the lower magnitude for aftershock detection, one can identify c with the time such that m c (c) = m a. According to this interpretation, the c obtained from experimental catalogs is a quantity caused by technical problems and does not contain any physical information. We wish to stress that, even if a deficit of small events at the beginning of a seismic sequence leads to longer c values, this cannot be responsible for the observed positive correlations between b and c. Indeed, let us indicate with b 0 and c 0 the true values of b and c, respectively, for a complete data set and with b m and c m their measured values affected by incompleteness. It is simple to show that STAI not only leads to c m > c 0, as explained above, but also to an average value of m larger than the true one producing a measured b m ( m m th 0.05) 1 systematically smaller than b 0. As a consequence, for fixed b 0 and c 0, STAI produces c m and b m that are inversely correlated, as opposed to the trend observed in experimental data. To confirm this interpretation we have generated numerically aftershock sequences by means of the ETAS model following the standard procedure outlined in Sec.II and assuming fixed values b 0 = 1.1, p = 1.1 and c 0 = 2500 sec. Each sequence contains on average 25000 aftershocks. To simulate the effect of incompleteness, we remove from the simulated catalogs aftershocks with magnitude m < m c ( t) according to Eq.(4). We set δ = 0.75 and consider different choices for m 1. Smaller values of m 1 lead to larger incompleteness and, as expected, produce larger c and smaller b values. The parametric plot c vs b obtained for different values of m 1 [0.5,2] is represented in Suppl. Fig.2. The functional dependence of c vs b is opposite to the one experimentally observed. To further support that the correlation between c and b is a physical effect, not related to spurious incompleteness, we perform the same study as in Fig.2 of the manuscript for different values of the aftershock thresholds m a = m M δ M. Incompleteness

5 2,0 10 4 average c value (sec) 1,5 10 4 1,0 10 4 5,0 10 3 0,0 0,8 1 1,2 b value δm=5 δm=4.5 δm=4 SC catalog ETAS FIG. 2. Parametric plot of thecvalue as function ofb. Black pluses are for mainshocks with magnitude in the range[2.5 : 4.5] in Southern California. Open symbols are for aftershocks with magnitude m m M δm, triggered by large mainshocks averaged over other geographic areas, for different values of δm: δm = 4 (green triangles), δm = 4.5 (red squares) and δm = 5 (blue circles). Filled magenta diamonds are results of numerical simulations of the ETAS model with STAI artificially implemented. is expected to be more pronounced for largerδ M. Conversely, Suppl. Fig.2 indicates that the linear relationship betweencandb weakly depends onδ M and therefore it is weakly affected by STAI. The same analysis (Suppl. Fig.3) for the parametric plotcvs L a for different choices of δ M also supports the correlation between c and L a as a physical feature of experimental aftershock sequences.

6 2,0 10 4 average c value (sec) 1,5 10 4 1,0 10 4 5,0 10 3 δm=5 δm=4.5 δm=4 SC catalog 0,0 0,2 0,4 0,6 L a FIG. 3. Parametric plot of thecvalue vs the size of the aftershock areal a. Black pluses are for mainshocks with magnitude in the range[2.5 : 4.5] in Southern California. Open symbols are results for aftershocks with magnitude m m M δm, triggered by large mainshocks averaged over other geographic areas, for different values of δm: δm = 4 (green triangles),δm = 4.5 (red squares) and δm = 5 (blue circles). IV. MAGNITUDE-FREQUENCY DISTRIBUTION AND AFTERSHOCKS IN THE NUMERICAL MODEL UNDER TECTONIC DRIVE In supplem. Fig.4 we plot results of numerical simulations of the model described in Fig.3 of the manuscript.

7 N(m) magnitude 3,5 4 4,5 5 5,5 6 6,5 7 7,5 10 4 10 3 10 2 10 1 10 0 7 10-1.1m magnitude 6 5 4 0 1e+08 2e+08 3e+08 4e+08 time FIG. 4. (Upper panel) The magnitude-frequency distribution for the simulated catalog is compared with the GR law of experimental catalog with b = 1.1 (orange broken line). (Lower Panel) A typical temporal window within the seismic history generated by the model of Fig.3 of the manuscript starting from a random initial condition. Time is expressed in seconds. After a transient regime of about 2.E8 sec the system exhibits main-aftershock sequences with spatio-temporal patterns very similar to experimental ones. Mainshocks are indicated by red vertical arrows. V. THE ROLE OF THE DYNAMIC FRICTION COEFFICIENT In supplem. Fig.5 we plot the number of aftershocks as function of the time from the mainshock for different values of the dynamic friction coefficient µ D [0,0.4]. Results indicate that the c-value, as well as the p-value, are not affected by µ D. The only effect of a larger µ D is a larger level of the background rate which corresponds to a flat behavior of the aftershock number at large time distances. The above result appears in agreement with experiments performed at high velocity suggesting that, for very low dynamic friction, significant weakening only occurs when sufficient displacement is accumulated. [1] P. Shearer, E. Hauksson, G. Lin, Bull. Seismol. Soc. Am. 95, 904 (2005). [2] F. Waldhauser, D. P. Schaff, J. Geophys. Res.: Solid Earth 113, B08311 (2008). [3] N.I.E.D., National Research Institute for Earth Science and Disaster Prevention (2012). [4] A.E.I.C., Alaska Earthquake Information Center, Catalog is available athttp : //www.aeic.alaska.edu/html d ocs/db2catalog.html (2011). [5] I.S.I.D.E., Italian Seismological Instrumental and Parametric Data-Base, Catalog is available at http : //iside.rm.ingv.it/iside/standard/ (2012). [6] T. Utsu, A. Seki, J. Seismol. Soc. Jpn. 7, 223 (1954). [7] Y. Ogata, Ann. Inst. Math.Statist. 50, 379 (1998). [8] A. Helmstetter, D. Sornette, Phys. Rev. E 66, 061104 (2002). [9] M. Bottiglieri, L. de Arcangelis, C. Godano, E. Lippiello, Phys. Rev. Lett. 104, 158501 (2010). [10] Y. Y. Kagan, Bulletin of the Seismological Society of America 94, 1207 (2004). [11] A. Helmstetter, Y. Kagan, D. Jackson, Bull. Seism. Soc. Am. 96(1), 90 (2006).

8 10 0 10-1 n(t-t 0 ) 10-2 (t-t 0 ) -1.1 10-3 µ D =0 µ D =0.2 µ D =0.4 10-4 10-1 10 0 10 1 10 2 10 3 t-t 0 (days) FIG. 5. The number of aftershocks as function of the temporal distance from the mainshock for different values of the friction coefficient µ D.