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

Similar documents
Long-Range Triggered Earthquakes That Continue After the Wavetrain Passes

Extending the magnitude range of seismic reservoir monitoring by Utilizing Hybrid Surface Downhole Seismic Networks

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

Delayed triggering of microearthquakes by multiple surface waves circling the Earth

Limitations of Earthquake Triggering Models*

New constraints on mechanisms of remotely triggered seismicity at Long Valley Caldera

High frequency identification of non volcanic tremor triggered by regional earthquakes

Periodically-triggered seismicity at Mt. Wrangell volcano following the Sumatra-Andaman Islands earthquake

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

Remotely Triggered Seismicity on the United States West Coast following the M w 7.9 Denali Fault Earthquake

Quantifying the remote triggering capabilities of large earthquakes using data from the ANZA Seismic Network catalog (southern California)

New constraints on mechanisms of remotely triggered seismicity at Long Valley Caldera

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

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

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

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

The Absence of Remotely Triggered Seismicity in Japan

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

Enhanced remote earthquake triggering at fluid injection sites in the Midwestern US

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

Dynamic Triggering Semi-Volcanic Tremor in Japanese Volcanic Region by The 2016 Mw 7.0 Kumamoto Earthquake

Remotely induced seismicity and the effects of regional earthquakes on open-vent systems

Interactions between earthquakes and volcano activity

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

Article in Proof. 2. Data and Method LXXXXX

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

The Coso Geothermal Area: A Laboratory for Advanced MEQ Studies for Geothermal Monitoring

Mechanics of Earthquakes and Faulting

Empirical Green s Function Analysis of the Wells, Nevada, Earthquake Source

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

Does Aftershock Duration Scale With Mainshock Size?

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

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

Automatic Moment Tensor Analyses, In-Situ Stress Estimation and Temporal Stress Changes at The Geysers EGS Demonstration Project

Distribution of volcanic earthquake recurrence intervals

Periodically Triggered Seismicity at Mount Wrangell, Alaska, After the Sumatra Earthquake

Earthquake Triggering at Alaskan Volcanoes Following the 3 November 2002 Denali Fault Earthquake

Applied Geophysics, Peking University, Beijing, China, 3 U.S. Geological Survey, Pasadena, California, USA

Effects of Surface Geology on Seismic Motion

MIGRATING SWARMS OF BRITTLE-FAILURE EARTHQUAKES IN THE LOWER CRUST BENEATH MAMMOTH MOUNTAIN, CALIFORNIA

DYNAMICALLY TRIGGERED EARTHQUAKES AND TREMOR: A STUDY OF WESTERN NORTH AMERICA USING TWO RECENT LARGE MAGNITUDE EVENTS. A Thesis.

Application of Phase Matched Filtering on Surface Waves for Regional Moment Tensor Analysis Andrea Chiang a and G. Eli Baker b

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

Determining the Earthquake Epicenter: Japan

DEVELOPMENT OF AUTOMATED MOMENT TENSOR SOFTWARE AT THE PROTOTYPE INTERNATIONAL DATA CENTER

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

Fracture induced shear wave splitting in a source area of triggered seismicity by the Tohoku-oki earthquake in northeastern Japan.

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

Testing the stress shadow hypothesis

FULL MOMENT TENSOR ANALYSIS USING FIRST MOTION DATA AT THE GEYSERS GEOTHERMAL FIELD

Some aspects of seismic tomography

Negative repeating doublets in an aftershock sequence

Earthquakes and Faulting

Remote triggered seismicity caused by the 2011, M9.0 Tohoku-Oki, Japan earthquake

Rupture Characteristics of Major and Great (M w 7.0) Megathrust Earthquakes from : 1. Source Parameter Scaling Relationships

A GLOBAL MODEL FOR AFTERSHOCK BEHAVIOUR

Module 7: Plate Tectonics and Earth's Structure Topic 4 Content : Earthquakes Presentation Notes. Earthquakes

Routine Estimation of Earthquake Source Complexity: the 18 October 1992 Colombian Earthquake

Lessons from (triggered) tremor

An autocorrelation method to detect low frequency earthquakes within tremor

to: Interseismic strain accumulation and the earthquake potential on the southern San

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

arxiv:physics/ v1 6 Aug 2006

IMPLEMENT ROUTINE AND RAPID EARTHQUAKE MOMENT-TENSOR DETERMINATION AT THE NEIC USING REGIONAL ANSS WAVEFORMS

The magnitude distribution of earthquakes near Southern California faults

Theory of earthquake recurrence times

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

Comparison of direct and coda wave stress drop measurements for the Wells, Nevada, earthquake sequence

Gutenberg-Richter Relationship: Magnitude vs. frequency of occurrence

Mechanical origin of aftershocks: Supplementary Information

Moment tensor inversion of near source seismograms

Spatial and temporal stress drop variations in small earthquakes near Parkfield, California

The Role of Asperities in Aftershocks

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

Aftershock From Wikipedia, the free encyclopedia

California foreshock sequences suggest aseismic triggering process

Plate Tectonics and Earth s Structure

Remote Triggering Not Evident Near Epicenters of Impending Great Earthquakes

Spatial clustering and repeating of seismic events observed along the 1976 Tangshan fault, north China

CAP M S Wallace. Vol. 27 No. 2 Jun EARTHQUAKE RESEARCH IN CHINA M S 4. 8 CAP. 3km - - P315

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

Masahiro Kosuga. Kosuga Earth, Planets and Space 2014, 66:77

Earthquakes.

Updated Graizer-Kalkan GMPEs (GK13) Southwestern U.S. Ground Motion Characterization SSHAC Level 3 Workshop 2 Berkeley, CA October 23, 2013

Source Duration Scales with Magnitude Differently For Earthquakes on the San Andreas Fault and on Secondary Faults in Parkfield, CA

Source of the July 2006 West Java tsunami estimated from tide gauge records

Earthquake patterns in the Flinders Ranges - Temporary network , preliminary results

Secondary Aftershocks and Their Importance for Aftershock Forecasting

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

Remote triggering of tremor along the San Andreas Fault in central California

Journal of Asian Earth Sciences

Testing for Poisson Behavior

Migrating swarms of brittle failure earthquakes in the lower crust beneath Mammoth Mountain, California

27th Seismic Research Review: Ground-Based Nuclear Explosion Monitoring Technologies

Earthquake Stress Drops in Southern California

Testing the stress shadow hypothesis

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

LETTERS. Non-volcanic tremor driven by large transient shear stresses

GEOPHYSICAL RESEARCH LETTERS, VOL. 34, LXXXXX, doi: /2007gl031077, 2007

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

Transcription:

Click Here for Full Article GEOPHYSICAL RESEARCH LETTERS, VOL. 33, L15313, doi:10.1029/2006gl026605, 2006 Long-range triggered earthquakes that continue after the wave train passes Emily E. Brodsky 1 Received 17 April 2006; revised 5 June 2006; accepted 27 June 2006; published 10 August 2006. [1] Large earthquakes can trigger distant earthquakes in geothermal areas. Some triggered earthquakes happen while the surface waves pass through a site, but others occur hours or even days later. Does this prolonged seismicity require a special mechanism to store the stress from the seismic waves that differs from ordinary aftershock mechanisms? These questions have driven studies of long-range triggering since the phenomenon s discovery. Here I attempt to answer the questions by examining the statistics of triggered sequences. Two separate observations are consistent with the prolonged sequences being simply local aftershocks of earthquakes triggered early in the wave train. First, the sequences obey Omori s Law over both short (1 hour) and longer (5 day) time intervals. Secondly, the number of observed triggered earthquakes in the first hour after the wave train can be predicted from the number of earthquakes triggered during the wave train. Even the very vigorous 10-day triggering at Long Valley from the 1992 Landers M w 7.3 earthquakes can be interpreted as the aftershocks of either a local M 4.1 earthquake or an equivalent combination of several smaller mainshocks. Therefore, long-range triggering does not need to include a mechanism to produce sustained stresses other than the process that generates aftershocks of the earthquakes that occur while the wave train is passing. Citation: Brodsky, E. E. (2006), Long-range triggered earthquakes that continue after the wave train passes, Geophys. Res. Lett., 33, L15313, doi:10.1029/2006gl026605. 1. Introduction [2] Seismic waves trigger earthquakes. For instance, as surface waves passed through geothermal sites like Yellowstone and Long Valley after earthquakes like the 1992 M w 7.3 Landers and 2002 M w 7.9 Denali events, local earthquakes flared up [Hill et al., 1993; Stark and Davis, 1996; Prejean et al., 2004]. However, the increased seismicity did not end with the end of the surface waves. It continued for an hour in some places and in other places for over a week. How do the transient waves of the far-field earthquakes make such a sustained response? This sustained response has been taken as a key piece of evidence for or against various triggering mechanisms [Gomberg, 2001; Brodsky et al., 1998]. Others have suggested that a handful of earthquakes are triggered locally during the passage of the seismic waves and all subsequent earthquakes are aftershocks of these first, locally triggered events [Hough and 1 Department of Earth Sciences, University of California, Santa Cruz, California, USA. Copyright 2006 by the American Geophysical Union. 0094-8276/06/2006GL026605$05.00 Kanamori, 2002]. Figure 1 illustrates the problem. Are the late triggered earthquakes aftershocks of the waveform triggered ones? [3] Here I show that in the best-documented cases, the sustained seismicity from long-range triggering (>100 km from the mainshock) in geothermal areas has the same behavior as ordinary aftershock sequences. The data will show that: (1) Long-range triggered sequences follow Omori s law and (2) The number of events triggered over 1 hour after the mainshock is consistent with number of aftershocks expected from earthquakes triggered during the passage of the seismic waves. Taken together, the data is fully consistent with the sustained long-range triggering occurring by the same process that sustains aftershock sequences in general. No more complicated process is required by this data set. 2. Omori s Law [4] Local earthquakes are poorly cataloged after a large earthquake, so I use broadband waveforms from individual stations to count the number of earthquakes near the station immediately after a main shock (Figure 2). [5] Figure 2 shows all documented long-range triggering sequences with more than 35 local events recorded on a broadband seismometers closer than 30 km away during the first hour after the arrival of the surface waves. There are only a handful of these cases. Two records come from a recent deployment of two Guralp 6TD seismometers (30 s corner frequency) in The Geysers geothermal field in Northern California during the 6/25/2005 Mendocino M w 7.2 earthquake and the third is from a permanent station. [6] A straight line on Figure 2 indicates an Omori s law with p-value of 1; that is, the cumulative number of earthquakes is Z T K N ¼ T 0 t dt ¼ K0 log T þ C where T is the time after the earthquake origin time T 0 and K, K 0 and C are constants. [7] Each individual sequence in Figure 2 follows an Omori s law decay of the form of equation (1). The fits are quite good as demonstrated by the correlation coefficient r which exceeds 97% for all cases. The fit reduces the sum of the residuals by a factor of 2.4 for the Geysers relative to a linear-time fit (constant seismicity rate) and a factor of 1.8 for Long Valley. Oddly, the fit is very good even during the first 400 s of the Denali-Long Valley sequence while the seismic wave excitation is still continuing. [8] Longer duration triggered sequences are in Figure 3. The identification of these sequences as triggered earth- ð1þ L15313 1of5

Figure 1. Vertical seismogram from the 2005 Mendocino earthquake recorded with a temporary deployment of broadband seismometers in The Geysers: (top) raw recording and (bottom) high-pass filtered at 7 Hz to show locally triggered earthquakes. The focus of this paper is whether or not the earthquakes labeled late triggering are aftershocks of those labeled waveform triggering. quakes is discussed elsewhere [Hill et al., 1993; Prejean et al., 2004]. All sequences with 100 cataloged earthquakes within 40 km and 5 days are included. The instantaneous rate is estimated using the nearest neighbor method which produces one rate measurement for each earthquake [Silverman, 1986]. Data is fit with a modified Omori s Law of the Omori s Law effect. For the Landers-Geysers and Denali-Yellowstone sequences, the best-fit values of p are 1.03 ± 0.03 and 0.98 ± 0.07, respectively, with significance >99%. This result is consistent with the earlier work of Husen et al. [2004] who fit the same Yellowstone data from the time of the Denali earthquake through 30 days with p = 1.02 ± 0.07 using a different statistical method. [9] The Landers-Long Valley sequence is more complex. Prior to 0.4 days, the data appears to be incomplete at low magnitudes (Figure 4) possibly due to the ongoing Landers aftershock sequence. If there is a variable completeness threshold in time, then early and late data should separately follow Omori s Law with the same value of p but different values of K. This is the case. Separately fitting the data for t < 0.2 days and t > 0.4 days, yields values of p of 0.95 ± 0.17 and 0.96 ± 0.12 for each set, respectively. If the whole time period is taken together for M > 2, the sparse data is best fit with p = 1.16 ± 0.17. [10] In summary, both the short and long-term data are consistent with Omori s Law (Figures 2 and 3). In some cases, the Omori s Law trend begins while the shaking was still occurring. This rate decay suggests that each entire prolonged sequence may be simply an aftershock sequence of mainshocks that occurred locally during the shaking. 3. Productivity [11] If the late-triggered earthquakes are aftershocks of the waveform-triggered ones, then the number of late- dn=dt ¼ K=t p ð2þ for times from the origin time of the earthquakes to when the seismicity rate reaches 50% of the average background rate. Once the aftershock rate approaches the background rate, other, non-related sequences are expected to overcome Figure 2. Cumulative number of earthquakes as counted from waveforms for broadband deployments near (<30 km) robust (>35 earthquakes) triggered sequences in the first hour after the arrival of the surface waves. Time 0 is the time of the arrival of the surface waves. Stars indicate the end of the mainshock wave trains. Geysers seismic data is from temporary deployments of two Guralp CMG-6TD seismometers and Long Valley data is from Univ. of Nevada broadband seismic station OMM. Figure 3. Seismicity rate of sequences with >100 cataloged earthquakes within 5 days. Solid lines show Omori s Law with best-fit values (equation (2)). (For Landers-Long Valley, the best-fit is on the >0.4 day data.) Vertical lines show times at which the seismicity rate reaches 50% of background rate for the sequence of the corresponding color. Background rates are simple averages of seismicity at each site in the years bracketing the triggering earthquake. The dotted line separates the Landers-Long Valley data into times of differing completeness as discussed in the text. Earthquake catalog data is from the Advanced National Seismic System (ANSS) catalog for Long Valley and Yellowstone and from the local Unocal catalog for The Geysers. Seismicity rate is smoothed with a 2-point running average. Only earthquakes above the completeness threshold are included. Cusps in the seismicity rate are secondary aftershock sequences. 2of5

Figure 4. Cumulative magnitude-frequency distribution for Long Valley earthquakes for time intervals following the Landers earthquake. The data does not follow a typical Gutenberg-Richter distribution until 0.5 days after Landers which suggests that the early time data may be incomplete for 1 < M <2. triggered earthquakes should be a function of the magnitude of the waveform-triggered earthquakes. Given information about the usual productivity of aftershocks in a region and the magnitude of the earthquakes during the waveform, one should be able to predict the number of late-triggered earthquakes. Making this prediction requires a model for aftershock productivity as a function of mainshock magnitude as well as a site with excellent coverage for both calibrating the model and observing the wavetrain triggered sequences. Here I use a standard model of aftershock productivity calibrated with 11 years of data from The Geysers to perform this exercise. [12] On average, the number of aftershocks N observed on the local network produced by a local mainshock magnitude M is N ¼ C10 am ð3þ where C and a are constants [Felzer et al., 2004; Helmstetter et al., 2005]. Figure 5 shows the typical productivity of small mainshocks at the Geysers based on local short-period data over an 11 year period excluding the times of the regional earthquakes to be studied. The figure records the mean number of earthquakes above the completeness threshold (local Unocal magnitude 0.5) within 15 km and an hour following mainshocks of magnitude M. Mainshocks are defined as earthquakes that have no larger earthquake within 1 day prior. This primitive mainshock separation criterion may not recover the true, physical productivity of each aftershock sequence, but it does provide an empirical method to measure the apparent productivity of the catalog. The observable apparent productivity is adequate for the prediction of late-triggered earthquakes from the waveform-triggered mainshocks because: (1) Contamination from unrelated sequences is just as likely to occur at the arbitrary time of the remote triggering as during any other time in the catalog and (2) The values of productivity are only used in the magnitude calibration range, so trade-offs between a and C are unimportant in this application. The 3495 disjoint sequences used in Figure 5 follow equation (3). The best-fit values of a is 0.6 ± 0.02 and C is 0.2 ± 0.02 where the errors are the standard deviation of 1500 bootstrap trials. [13] Using the magnitudes of the earthquakes in a handpicked catalog for the 10 minutes following 9 regional earthquakes and the calibration in Figure 5, I predict the number of earthquakes in the next hour (Figure 6). Ten minutes covers the extended coda for the large regional events as can be seen in Figure 1. Each earthquake during the wavetrain is treated as a separate mainshock. The total predicted number of aftershocks is the sum of the contributions from each wavetrain triggered earthquake. [14] The observed number of events fits the predictions with 98% significance. Several cases have 10 20% underpredictions of the sustained seismicity which might be attributable to catalog incompleteness during the waveforms. The Figure 5. Number of earthquakes at The Geysers within one hour of mainshocks of magnitude M. Magnitudes are local Unocal network magnitudes which are typically depressed 0.5 units from ANSS. Figure 6. Predicted and observed number of earthquakes at The Geysers within 1 hour after distance mainshocks. Studied earthquakes were: 1992 M w 7.3 Landers, 1993 M w 6.0 Klamath Falls, 1994 M w 7.0 Mendocino, 1994 M w 6.6 Northridge, 1999 M w 7.1 Hector Mine, 2002 M w 7.8 Denali, 2003 M w 6.6 San Simeon, 2003 M w 8.3 Tokachi-Oki, 2005 M w 7.2 Mendocino. The number of predicted and observed earthquakes for Northridge and Klamath Falls are identical, so the points lie on top of each other in this figure. 3of5

outlier is the 1994 Mendocino earthquake which had fewer than expected earthquakes after the waves passed. At least 80% of the triggered earthquakes can be explained without any sustained process other than local aftershock generation. 4. Long Valley and Landers: A Counter-Example? [15] The most prolonged example of long-range triggering is Landers at Long Valley, which lasted 10 days. At first glance, this sequence may appear to be a counter-example to the hypothesis that prolonged triggering is a local aftershock sequence. Here I show that the statistics of the sequence do not contradict the hypothesis and the prolonged triggering is consistent with a locally triggered aftershock sequence. [16] As shown above, the sequence follows Omori s Law. No local seismometers remained operational and on-scale during the Landers waves, so the productivity prediction performed in Figure 6 is impossible for this data set. Instead, I perform the inverse exercise of predicting the magnitude of the largest local earthquakes during the wave train from the total number of earthquakes. For Long Valley from 1988 1999, I find a = 0.5 ± 0.05 and C = 1.8 ± 0.6 for M 1 earthquakes between 0.2 and 10 days after the mainshock (Main shocks are again defined as earthquakes with no larger earthquake 1 day prior; Errors are again 1 standard deviation on 1500 bootstraps). [17] The observed number of earthquakes (208) is consistent with the entire sequence being triggering by an M4.1 local earthquake. Alternatively, several smaller earthquakes during the wave train could have an equivalent effect. For instance, three M L 3.1 earthquakes would suffice. Propagating the errors on a and C yields a range for the equivalent mainshock magnitude of M = 3.5 5.0. [18] Magnitude 3 4 earthquakes in Long Valley could be hidden in the wave train of the Landers earthquake. The closest functioning broadband seismometer was 140 km away. The short-period seismometers were too saturated during the main part of the Rayleigh wave to show local events even with very aggressive filtering. Shaking during the Landers earthquake woke people in the Long Valley area and would not necessarily be discernible from shaking originating locally. [19] The observed deformation at Long Valley is also consistent with a local earthquake [Johnston et al., 1995]. Roeloffs et al. [2003] observed that the strainmeter record of deformation at Long Valley from Landers was nearly identical to deformation from a local M4.9 earthquake. [20] Other potential counter-examples are late large events that produce bursts of seismicity several days after the mainshock. Examples include the Denali-triggered earthquakes at Long Valley, Rainier and Yellowstone and the M L 5.6 Little Skull Mountain earthquake a day after the Landers earthquake [Prejean et al., 2004; Smith et al., 2001]. All of these sites had small earthquakes that began as soon as the seismic waves from the mainshock passed. Like all aftershock sequences, the continuing seismicity should have bursts of activity from large secondary aftershocks (See, for example, the cusps in Figure 3). The late bursts may be particularly large secondary aftershocks of the sequences that began with the passage of the surface waves. 5. Spatial Distribution [21] Locations are too poor to definitively constrain the location of late triggered earthquakes relative to early ones. At The Geysers, the earthquake location errors during the wave train are much higher than the usual 0.7 km average horizontal error. 6. Conclusion [22] For the analyzed sequences, the late triggered events appear to be aftershocks of the earthquakes triggered during the passage of the seismic waves. No sustained stress beyond the usual aftershock process is required. More complicated processes might occur, but there is nothing in the data that requires them. [23] Transient triggering by the dynamic stresses of the Rayleigh waves followed by local static stress triggering of aftershocks would be consistent with the data presented here. However, other studies [Felzer and Brodsky, 2006] indicate that aftershocks may also be triggered by seismic waves. In that case, a sustained stress mechanism is still necessary for the aftershock process. Geothermal areas may be more susceptible to triggering from shaking than ordinary crust [Brodsky et al., 2003; Manga and Brodsky, 2006], but once started, the triggered sequence proceeds like any other series of aftershocks. The conclusion of this study is that the prolonged nature of long-range triggered sequences is no more (or less) mysterious than the prolonged nature of aftershock sequences in general. [24] Acknowledgments. Many thanks to R. Haught, W. Mooney, R. Sell and M. Stark for making the field deployments happen. M. Walters picked the Geysers short-period records. Additional catalog and waveform data is from ANSS, NCEDC and IRIS. The work was supported in part by NSF. Constructive reviews from J. Gomberg and D. Marsan improved the paper. References Brodsky, E. E., B. Sturtevant, and H. Kanamori (1998), Earthquake, volcanoes and rectified diffusion, J. Geophys. Res., 103, 23,827 23,838. Brodsky, E. E., E. Roeloffs, D. Woodcock, I. Gall, and M. Manga (2003), A mechanism for sustained groundwater pressure changes induced by distant earthquakes, J. Geophys. Res., 108(B8), 2390, doi:10.1029/ 2002JB002321. Felzer, K. R., and E. E. Brodsky (2006), Evidence for dynamic aftershock triggering from earthquake densities, Nature, 441, 735 738. Felzer, K., R. Abercrombie, and G. Ekstrom (2004), A common origin for aftershocks, foreshocks, and multiplets, Bull. Seismol. Soc. Am., 94, 88 98. Gomberg, J. (2001), The failure of earthquake failure models, J. Geophys. Res., 106, 16,253 16,263. Helmstetter, A., Y. Y. Kagan, and D. D. Jackson (2005), Importance of small earthquakes for stress transfers and earthquake triggering, J. Geophys. Res., 110, B05S08, doi:10.1029/2004jb003286. Hill, D. P., et al. (1993), Seismicity remotely triggered by the magnitude 7.3 Landers, California, earthquake, Science, 260, 1617 1623. Hough, S., and H. Kanamori (2002), Source properties of earthquakes near the Salton Sea triggered by the 16 October 1999 M 7.1 Hector Mine, California, earthquake, Bull. Seismol. Soc. Am., 92, 1281 1289. Husen, S., S. Wiemer, and R. Smith (2004), Remotely triggered seismicity in the Yellowstone National Park region by the 2002 Mw = 7.9 Denali Fault earthquake, Alaska, Bull. Seismol. Soc. Am., 94, S317 S331. Johnston, M. J. S., D. P. Hill, A. T. Linde, J. Langbein, and R. Bilham (1995), Transient deformation during triggered seismicity from the 28 June 1992 M w = 7.3 Landers earthquake at Long Valley volcanic caldera, California, Bull. Seismol. Soc. Am., 85, 787 795. 4of5

Manga, M., and E. E. Brodsky (2006), Seismic triggering of eruptions in the far field: Volcanoes and geysers, Annu. Rev. Earth Planet. Sci., 34, 263 291. Prejean, S. G., et al. (2004), Observations of remotely triggered seismicity on the United States West Coast following the M7.9 Denali Fault earthquake, Bull. Seismol. Soc. Am., 94, S348 S359. Roeloffs, E., M. Sneed, D. L. Galloway, M. Storey, C. D. Farrar, J. Howle, and J. Hughes (2003), Water-level changes induced by local and distant earthquakes at Long Valley caldera, California, J. Volcanol. Geotherm. Res., 127(3 4), 269 303. Silverman, B. (1986), Density Estimation for Statistics and Data Analysis, CRC Press, Boca Raton, Fla. Smith, K. D., J. N. Brune, D. depolo, M. K. Savage, R. Anooshehpoor, and A. F. Sheehan (2001), The 1992 Little Skull Mountain earthquake sequence, southern Nevada Test Site, Bull. Seismol. Soc. Am., 91, 1595 1606. Stark, M. A., and S. D. Davis (1996), Remotely triggered microearthquakes at The Geysers geothermal field, California, Geophys. Res. Let., 23, 945 948. E. E. Brodsky, Department of Earth Sci., UC Santa Cruz, Santa Cruz, CA 95060, USA. (brodsky@es.ucsc.edu) 5of5