Constraints from precariously balanced rocks on preferred rupture directions for large earthquakes on the southern San Andreas Fault

Similar documents
RUPTURE MODELS AND GROUND MOTION FOR SHAKEOUT AND OTHER SOUTHERN SAN ANDREAS FAULT SCENARIOS

ShakeOut-D: Ground motion estimates using an ensemble of large earthquakes on the southern San Andreas fault with spontaneous rupture propagation

TeraShake2: Spontaneous Rupture Simulations of M w 7:7 Earthquakes on the Southern San Andreas Fault

Precarious Rocks Methodology and Preliminary Results

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

VALIDATION AGAINST NGA EMPIRICAL MODEL OF SIMULATED MOTIONS FOR M7.8 RUPTURE OF SAN ANDREAS FAULT

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

Deterministic Generation of Broadband Ground Motions! with Simulations of Dynamic Ruptures on Rough Faults! for Physics-Based Seismic Hazard Analysis

SDSU Module Kim Olsen and Rumi Takedatsu San Diego State University

Effects of Surface Geology on Seismic Motion

3D VISCO-ELASTIC WAVE PROPAGATION IN THE BORREGO VALLEY, CALIFORNIA

ACCOUNTING FOR SITE EFFECTS IN PROBABILISTIC SEISMIC HAZARD ANALYSIS: OVERVIEW OF THE SCEC PHASE III REPORT

Eleventh U.S. National Conference on Earthquake Engineering Integrating Science, Engineering & Policy June 25-29, 2018 Los Angeles, California

Synthetic Seismicity Models of Multiple Interacting Faults

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

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

Effects of 3D basin structure on long-period ground motions in SW British Columbia, Canada, for large scenario earthquakes

Simulation of earthquake rupture process and strong ground motion

Bulletin of the Seismological Society of America, Vol. 94, No. 6, pp , December 2004

GROUND MOTION TIME HISTORIES FOR THE VAN NUYS BUILDING

Short Note Source Mechanism and Rupture Directivity of the 18 May 2009 M W 4.6 Inglewood, California, Earthquake

Lower Bounds on Ground Motion at Point Reyes During the 1906 San Francisco Earthquake from Train Toppling Analysis

GROUND MOTION TIME HISTORIES FOR THE VAN NUYS BUILDING

EARTHQUAKE LOCATIONS INDICATE PLATE BOUNDARIES EARTHQUAKE MECHANISMS SHOW MOTION

SCEC INCITE OLCF PROJECT PROGRESS REPORT,

GROUNDWORK FOR USING PRECARIOUSLY BALANCED ROCKS TO CONSTRAIN SEISMIC HAZARD MODELS IN NEW ZEALAND

Directivity of near-fault ground motion generated by thrust-fault earthquake: a case study of the 1999 M w 7.6 Chi-Chi earthquake

EFFECTS OF CORRELATION OF SOURCE PARAMETERS ON GROUND MOTION ESTIMATES

Effects of Surface Geology on Seismic Motion

Jack Loveless Department of Geosciences Smith College

Topography on Earthquake Motions in Sedimentary Basins

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

SIMULATING STRONG GROUND MOTION FROM COMPLEX SOURCES BY RECIPROCAL GREEN FUNCTIONS ABSTRACT

Modelling Strong Ground Motions for Subduction Events in the Wellington Region, New Zealand

The SDSU Broadband Ground Motion Generation Module BBtoolbox Version 1.5

THREE-DIMENSIONAL FINITE DIFFERENCE SIMULATION OF LONG-PERIOD GROUND MOTION IN THE KANTO PLAIN, JAPAN

Near-Source Ground Motion along Strike-Slip Faults: Insights into Magnitude Saturation of PGV and PGA

Effects of Fault Dip and Slip Rake Angles on Near-Source Ground Motions: Why Rupture Directivity Was Minimal in the 1999 Chi-Chi, Taiwan, Earthquake

High-Frequency Ground Motion Simulation Using a Source- and Site-Specific Empirical Green s Function Approach

High Resolution Imaging of Fault Zone Properties

Motion ERIC M. JONES LOS ALAMOS NATIONAL LABORATORIES LOS ALAMOS, NEW MEXICO

Imaging sharp lateral velocity gradients using scattered waves on dense arrays: faults and basin edges

Effects of Surface Geology on Seismic Motion

CONTROLLING FACTORS OF STRONG GROUND MOTION PREDICTION FOR SCENARIO EARTHQUAKES

Seismic hazard analysis of Tianjin area based on strong ground motion prediction

Earthquake ground motion prediction using the ambient seismic field

EFFECTS OF SIMULATED MAGNITUDE 9 EARTHQUAKE MOTIONS ON STRUCTURES IN THE PACIFIC NORTHWEST

SCEC Broadband Platform (BBP) Simulation Methods Validation for NGA-East

Physics-Based 3D Ground Motion Simulations The SCEC/CME High-F and SEISM Projects!

GROUND MOTIONS FROM LARGE EARTHQUAKES (MW³7) ON THE SANTA MONICA MOUNTAIN THRUST AND HOLLYWOOD-SANTA MONICA-MALIBU FAULTS

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

SCENARIO MODELING OF THE 2014 Mw6.0 SOUTH NAPA, CALIFORNIA, EARTHQUAKE USING AN ADVANCED BROADBAND KINEMATIC SOURCE MODEL

Outstanding Problems. APOSTOLOS S. PAPAGEORGIOU University of Patras

CAMPBELL-BOZORGNIA NEXT GENERATION ATTENUATION (NGA) RELATIONS FOR PGA, PGV AND SPECTRAL ACCELERATION: A PROGRESS REPORT

State of Stress in Seismic Gaps Along the SanJacinto Fault

UPDATED GRAIZER-KALKAN GROUND- MOTION PREDICTION EQUATIONS FOR WESTERN UNITED STATES

Widespread Ground Motion Distribution Caused by Rupture Directivity during the 2015 Gorkha, Nepal Earthquake

Fault Specific, Dynamic Rupture Scenarios for Strong Ground Motion Prediction

Arthur Frankel, William Stephenson, David Carver, Jack Odum, Robert Williams, and Susan Rhea U.S. Geological Survey

3D waveform simlation in Kobe of the 1995 Hyogoken-Nanbu earthquake by FDM using with discontinuous grids

Seismic and aseismic processes in elastodynamic simulations of spontaneous fault slip

Earthquakes and Seismotectonics Chapter 5

Lab 6: Earthquake Focal Mechanisms (35 points)

CHARACTERIZING EARTHQUAKE SLIP MODELS FOR THE PREDICTION OF STRONG GROUND MOTION

CyberShake: A Physics-Based Seismic Hazard Model for Southern California

STRIKE SLIP SPLAY USING DYNAMIC RUPTURE MODELS

SOURCE MODELING OF RECENT LARGE INLAND CRUSTAL EARTHQUAKES IN JAPAN AND SOURCE CHARACTERIZATION FOR STRONG MOTION PREDICTION

Tests and Applications of 3-D Geophysical Model Assembly

Root-mean-square distance and effects of hanging wall/footwall. Wang Dong 1 and Xie Lili 1,2

Elastic rebound theory

Development of U. S. National Seismic Hazard Maps and Implementation in the International Building Code

Yesterday scaling laws. An important one for seismic hazard analysis

Mechanics of Earthquakes and Faulting

Low-velocity damaged structure of the San Andreas Fault at Parkfield from fault zone trapped waves

Figure Locations of the CWB free-field strong motion stations, the epicenter, and the surface fault of the 1999 Chi-Chi, Taiwan earthquake.

ATTENUATION RELATIONSHIP FOR ESTIMATION OF PEAK GROUND VERTICAL ACCELERATION USING DATA FROM STRONG MOTION ARRAYS IN INDIA

Knowledge of in-slab earthquakes needed to improve seismic hazard estimates for southwestern British Columbia

GROUND MOTION SPECTRAL INTENSITY PREDICTION WITH STOCHASTIC GREEN S FUNCTION METHOD FOR HYPOTHETICAL GREAT EARTHQUAKES ALONG THE NANKAI TROUGH, JAPAN

CHARACTERIZATION OF DIRECTIVITY EFFECTS OBSERVED DURING 1999 CHI-CHI, TAIWAN EARTHQUAKE

The Three-Dimensional Dynamics of a Nonplanar Thrust Fault

BROADBAND MODELING OF STRONG GROUND MOTIONS FOR PREDICTION PURPOSES FOR SUBDUCTION EARTHQUAKES OCCURRING IN THE COLIMA-JALISCO REGION OF MEXICO

2D versus 1D ground-motion modelling for the Friuli region, north-eastern Italy

Kinematics of the Southern California Fault System Constrained by GPS Measurements

Development of Procedures for the Rapid Estimation of Ground Shaking Task 7: Ground Motion Estimates for Emergency Response Final Report

Hemanth Siriki 1, Harsha S. Bhat 2,Xiao Lu 3,Swaminathan Krishnan 4. Abstract

Seismic Reflection Views of the 1994 Northridge Earthquake Hypocentral Region Using Aftershock Data and Imaging Techniques

RECIPE FOR PREDICTING STRONG GROUND MOTIONS FROM FUTURE LARGE INTRASLAB EARTHQUAKES

Dynamic source inversion for physical parameters controlling the 2017 Lesvos earthquake

BROADBAND STRONG MOTION SIMULATION OF THE 2004 NIIGATA- KEN CHUETSU EARTHQUAKE: SOURCE AND SITE EFFECTS

STUDYING THE IMPORTANT PARAMETERS IN EARTHQUAKE SIMULATION BASED ON STOCHASTIC FINITE FAULT MODELING

involved in multiple state-wide re-evaluations of seismic risk by the State of California.

March 2017 SCEC Rupture Dynamics Code Validation Workshop

Proposed modification on velocity attenuation relationships in West of Iran with special respect to Dorood fault

Characterizing Earthquake Rupture Models for the Prediction of Strong Ground Motion

Simulations in the Los Angeles Basin

Variability of Near-Field Ground Motion from Dynamic Earthquake Rupture Simulations

Equivalent Medium Parameters for Numerical Modeling in Media with Near-Surface Low Velocities

Earthquake stress drop estimates: What are they telling us?

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

Time-varying and long-term mean aftershock hazard in Wellington

Transcription:

DOI 10.1007/s10950-007-9078-7 ORIGINAL ARTICLE Constraints from precariously balanced rocks on preferred rupture directions for large earthquakes on the southern San Andreas Fault Kim Olsen & James Brune Received: 22 March 2007 / Accepted: 22 November 2007 # Springer Science + Business Media B.V. 2007 Abstract We have compared near-fault ground motions from TeraShake simulations of M w 7.7 earthquake scenarios on the southern San Andreas Fault with precariously balanced rock locations. The Tera- Shake scenarios with different directions of rupture generate radically different ground motions to the northwest of the Los Angeles Basin, primarily because of directivity effects, and thus provide constraints on the ground motion and rupture direction for the latest (1690) large event on that section of the San Andreas Fault. Due to the large directional near-field ground motions predicted by the simulations, we expect the precariously balanced rocks to be located primarily in the backward rupture direction or near the epicenter. Preliminary results favor persistent nucleation at or slightly northwest of the San Gorgonia Pass fault zone for large earthquakes on the southern San Andreas Fault. Keywords Finite-difference simulation. Long-period ground motion. Precariously balanced rocks. San Andreas Fault. Rupture directivity K. Olsen (*) Dept of Geological Sciences, San Diego State University, 5500 Campanile Dr, San Diego, CA 92182, USA e-mail: kbolsen@sciences.sdsu.edu J. Brune Nevada Seismological Laboratory, University of Reno, Reno, NV 89557, USA 1 Introduction The two segments of the San Andreas Fault south of the 1906 rupture, the San Bernardino Mountains segment and the Coachella Valley segment, have not produced major earthquakes since 1812 and about 1690 (Weldon et al. 2004), respectively. The average recurrence interval for large events with surface rupture on these segments are 146+91 60 and 220± 13 years, respectively (Working Group on California Earthquake Probabilities 1995). To estimate the expected ground motions in southern California for such event, Olsen et al. (2006, 2008) simulated 4 min of 0 0.5-Hz wave propagation for M w 7.7 earthquake scenarios between Cajon Creek and Bombay Beach (see Fig. 1), the TeraShake simulations. The areal extent of the TeraShake simulations was a rectangular region, 600 km along N50W and 300 km along N40E (see Fig. 1). The simulations used a 3,000 by 1,500 by 400 mesh, or 1.8 billion cubes with dimensions 200 m 3. The three-dimensional crustal structure was a subset of the Southern California Earthquake Center (SCEC) Community Velocity Model (Magistrale et al. 2000) Version 3.0, with elastic parameters constrained by gravity and reflection seismic data, oil-company drill holes, and shallow geotechnical borings. The depth extent of the model is 80 km. The near-surface S-wave velocity was truncated at 500 m/s due to computational limitations. However, comparisons with finite element synthetics computed using very fine discretization of the low-velocity layers suggest

that the neglect of velocities below 500 m/s in the SCEC model causes at most a small underestimate of amplitudes in the 0 0.5-Hz band, typically less than about 10 20%. No surface topography was included in the simulations. Two sets of simulations for large earthquakes on the southern San Andreas Fault were carried out, TeraShake-1 and TeraShake-2. For both sets of simulations, the San Andreas Fault geometry was approximated as five vertical, planar segments from the 2002 US Geological Survey (USGS) National Hazard Maps (Fig. 1). The rupture length is 200 km, and the down-dip width is 15 km. The TeraShake-1 simulations (Olsen et al. 2006) used a kinematic description of the source model based on that inferred for the 2002 M w 7.9 Denali earthquake (Oglesby et al. 2004). The source is modeled as a kinematic slip function with six 1-s time-windows of Brune pulses of 0.5-s rise time. This source model was found to produce 0.01 0.5-Hz synthetic seismograms in good agreement with observed strong motion data (Oglesby et al. 2004). The TeraShake-2 simulations (Olsen et al. 2008) used a more realistic mechanical (dynamic) description of the rupture propagation. A slip-weakening friction law with initial stress and slipweakening distance inferred from dynamic models of the 1992 Landers, CA, USA earthquake (Peyrat et al. 2001) was used in the dynamic rupture propagation. The TeraShake-1 (kinematic) and TeraShake-2 (dynamic) sources show important differences and similarities. The maximum slip and sliprate for both sets of rupture models reach about 10 m and 5 m/s, respectively. The kinematic source is characterized by relatively large, smooth asperities and a constant rupture velocity. The dynamic source is characterized by strong variations in rupture speed as well as frequent separation into several slipping areas of highly varying shape and sliprate. However, the average rupture velocity for the TeraShake-2 scenarios is close to the constant value of 3.3 km/s used for TeraShake-1, generating a rupture duration of approximately 70 s for both sets of rupture models. The duration of slip (at a given point) is similar in the kinematic and dynamic models, although much more variable in the latter model. Simulations were carried out for kinematic and dynamic sources initiating at the northwestern end rupturing toward the southeast (NW SE; TS1.2, Fig. 1 Location map for the TeraShake simulations. The red rectangle depicts the simulations area, which is rotated 40 clockwise from North. The black rectangle depicts the area used for comparison of PGVs and locations of PBRs. The dotted line depicts the part of the San Andreas Fault that ruptured in the TeraShake simulations

TS2.3), starting at the southeastern end and rupturing toward the northwest (SE NW; TS1.3 4, TS2.1 2) and for a bilateral rupture (TS1.5) with kinematic source. The peak ground velocity (PGV) distributions for both TeraShake-1 and TeraShake-2 simulations reveal a striking contrast in ground motion pattern between NW SE versus SE NW rupture scenarios. The ground motion is focused in the direction of rupture, and this rupture directivity effect is dramatically modified by interactions with the chain of sedimentary basins (the San Bernardino, Chino, San Gabriel, and Los Angeles basins) running westward from the northern terminus of the rupture to downtown Los Angeles. This chain of basins forms a lowvelocity structure that acts as a waveguide, trapping seismic energy along the southern edge of the San Bernardino and San Gabriel Mountains and channeling it into the Los Angeles region. This guided wave is efficiently excited by the SE NW rupture scenarios, but not appreciably by the NW SE rupture scenarios. In contrast, the NW SE scenarios would generate PGVs of about an order of magnitude smaller in the Los Angeles basin, but strong reverberations for an extended duration at locations inside the Salton Trough, including the Coachella and Imperial Valleys (Olsen et al. 2006, 2008). While the TeraShake-1 and TeraShake-2 simulations produce similar spatial patterns of ground motion, the dynamic sources decrease the maximum ground motions obtained by the kinematic source, in some areas of Los Angeles by factors of 2 3. The primary reason for the smaller ground motions is a more incoherent wavefield radiated from the complex dynamic source. This complexity consists mainly of abrupt changes in rupture speed and direction, which produces characteristic sun burst patterns of elevated ground motions along bands extending outward from the fault (Olsen et al. 2008). The TeraShake simulations were carried out using a 4th-order staggered-grid finite difference code (Olsen 1994) with a coarse-grained implementation of the memory variables for a constant-q solid (Day and Bradley 2001) and Q relations were validated against data (Olsen et al. 2003). The dynamic rupture propagation used in the TeraShake-2 simulations was implemented using the stress-glut method (Andrews 1999). We use the efficient Perfectly Matched Layers absorbing boundary conditions on the sides and bottom of the model (Marcinkovich and Olsen 2003). 2 Correlation of precariously balanced rocks and TeraShake PGVs In lack of strong ground motion recordings, the location of granitic precariously balanced rocks (PBRs) have been used to estimate upper bounds of the seismic shaking for periods sufficiently long to sample an earthquake rupture forecast (Bell et al. 2004; Purvance 2005; Purvance et al. 2005). PBRs are mapped near the section of the southern San Andreas Fault that ruptured in the TeraShake simulations (see Fig. 2) by James Brune and colleagues. Here, we use the PBRs to establish bounds on the intensity of shaking and compare to the PGV (here defined as the root sum of squares of all three components) patterns from TeraShake simulations with different rupture directions and source descriptions. A comparison of the locations of PBRs with TeraShake-1 and TeraShake-2 PGV distributions are shown in Figs. 3 and 4, with the 50-cm and 35-cm PGV contours superimposed for the TeraShake-1 and TeraShake-2 results, respectively. The relatively smaller reference PGV for TeraShake-2 is due to the smaller ground motions generated from the more complex dynamic sources, as discussed above. The PGV contours are somewhat arbitrarily selected Fig. 2 Precariously balanced rocks at Lovejoy Buttes

scenarios with rupture starting either at the NW or SE end of the fault segment (Fig. 3, TS1.2 4). In particular, the two PBRs with the smallest distance to the segments of the San Andreas Fault used in the TeraShake simulations, located near Banning and Beaumont, may provide important clues to a preferred rupture direction of the fault. These two PBRs tend to experience consistently large PGVs for unilateral rupture scenarios but much smaller PGVs for a simulation with bilateral rupture with nucleation near these two sites (TS1.5, see Fig. 3). Thus, the correlation of PBRs with TeraShake-1 PGVs for the available simulations tends to favor bilateral over unilateral rupture directions on the southern San Andreas Fault from the TeraShake-1 simulations. The correlation of PBRs with PGVs from Tera- Shake-2 simulations (Fig. 4) provides less guidance for the rupture direction, in part due to the lack of a Fig. 3 PGV distributions from four TeraShake-1 simulations (TS1.2 5). The dashed line is the section of the San Andreas Fault that ruptured in the simulations. The solid line is the 50- cm/s PGV contour. The triangles depict precarious rock locations, with sites at Banning and Beaumont in magenta and Lovejoy Buttes in yellow because peak ground accelerations (PGAs) are required to determine absolute toppling thresholds for the PBRs. However, the shape of the PGV contours may provide guidance to a preferred rupture direction on the southern San Andreas Fault. Note the characteristic cone-shaped pattern for the PGVs, with a relatively narrow width along the fault around the epicenter and widening along the fault away from the epicenter. Several PBRs are located at sites experiencing ~50 cm/s or higher PGV for the TeraShake-1 Fig. 4 PGV distributions from four TeraShake-2 simulations (TS2.1 3). The dashed line is the section of the San Andreas Fault that ruptured in the simulations. The solid line is the 35- cm/s PGV contour. The triangles depict precarious rock locations, with sites at Banning and Beaumont in magenta and Lovejoy Buttes in yellow

bilateral rupture scenario with dynamic source description. In addition, the stronger local variation of the PGVs for TeraShake-2 due to the larger complexity of the dynamic sources complicates the correlation with the PBRs. However, similar cone-shaped PGV patterns as noted for the TeraShake-1 simulations prevail for the dynamic sources. The TS2.3 simulation (rupture toward the SE) generates, on average, a narrower cone of PGVs expanding in the rupture direction for a M w 7.7 event occurring on this part of the San Andreas Fault, with all mapped PBRs experiencing PGVs less than 35 cm/s. In contrast, several PBRs are located inside or near (including those at Banning and Beaumont) the 35-cm/s contour for TS2.1 2 (rupture toward the NW). Thus, the TeraShake-2 results tend to favor a NW SE rupture direction over a SE NW rupture direction. 3 Discussion and conclusions We have used comparison of near-fault ground motions from four TeraShake-1 and three Tera- Shake-2 scenario earthquake simulations with PBR locations to constrain the preferred rupture direction of the southern San Andreas Fault. The observed PBRs closest to the segments of the San Andreas Fault rupturing in the TeraShake simulations are located near Banning and Beaumont, approximately 1/3 of the distance from the NW end to the SE end of the rupturing fault. The presence of the two near-fault PBRs suggests that large earthquakes on the southern San Andreas consistently generate lower strong ground motions along the fault in this area. Analysis of near-fault rupture patterns for the TeraShake simulations suggests that the epicentral area generally experiences smaller PGVs as compared to sites along the fault in the rupture direction. This observation is consistent with the effects of rupture directivity for large, shallow-crustal earthquakes on near-vertical faults. Furthermore, the seven TeraShake simulations cover a realistic range of maximum slip and sliprates, including the part of the fault near the Banning and Beaumont PBRs (see Fig. 5). Thus, any bias from the rupture models affecting the correlation of PGVs and PBRs is likely limited. We conclude that the distribution of PBRs is consistent with persistent nucleation of large earthquakes on the southern San Andreas Fault near Palm Springs (see Fig. 1). The PGVs for the bilateral rupture scenario TS1.5, with epicenter close to the Banning and Beaumont PBRs, and the NW SE rupture scenario TS2.3, produce the best correlation with the location of near-fault PBRs. Two different possibilities for a preferred rupture direction, both with nucleation toward the north central or northern part of the causative fault in the TeraShake scenarios, are consistent with the results: (1) that successive ruptures on the southern San Andreas have propagated both directions, or (2) that the ruptures propagate bilaterally. The geometry of the San Andreas Fault shows increased complexity in the proposed area of persistent nucleation. The San Andreas Fault is divided into two sub-parallel traces, the Mission Creek fault and the Banning fault near this area. Here, the NW SE strike of the San Andreas Fault to the NW and SE is replaced by a WNW ESE strike in the San Gorgonio fault zone. The M w 6.1 1986 North Palm Springs and 1948 M L 6.3 Desert Hot Springs earthquakes occurred in this area, with epicenters between the surface traces of the Banning and Mission Creek faults. The focal mechanisms for both events occurred on NE-dipping faults with a thrust component (e.g., Nicholson 1996), in agreement with the significant vertical relief of the mountains in this area. The stress conditions required for nucleation of an earthquake depends on the geometry of the fault (e.g., King et al. 1994), and it is possible that the regional stress accumulation may be more favorable for nucleation on a dipping fault rather than a near-vertical fault geometry (as used for the model for the San Andreas Fault in this study). An example of such scenario was the 2002 M w 7.9 Denali, AK, USA earthquake, which was initiated by a thrust earthquake at one end. Moreover, Freed and Lin (2002) showed that the nearby 1992 M w 7.3 Landers and 1999 M w 7.1 Hector Mine earthquakes brought the San Andreas Fault closer to failure in the San Gorgonio Pass area. Finally, it should be noted that the concentration of PBRs near Banning and Beaumont might be explained by reduced ground motions on the footwall of San Gorgonio Pass fault zone, due to the hanging wall effect (Abrahamson and Somerville 1996). The TeraShake simulations currently only contain frequencies less than 0.5 Hz, while near-fault peak motions oftentimes occur at frequencies larger than 0.5 Hz. Thus, near-fault peak ground motions estimated from the simulations, in particular PGAs

Fig. 5 (top) Maximum slip and (bottom) maximum sliprate for the TeraShake simulations from the NW (left) to the SE (right) end of the segments of the San Andreas fault rupturing in the TeraShake simulations (dotted line in Fig. 1) required to determine the toppling threshold for the PBRs, are likely underestimated. Future efforts should therefore work toward increasing the maximum frequency of the synthetic seismograms for the southern San Andreas scenarios. Due to computational limitations, the largest frequency for deterministic ground motion models is currently limited to 1 2 Hz, when realistic near-surface velocities are included (see, for example, Benites and Olsen 2005). Instead, several recent methods have been proposed to generate realizations of broadband synthetic seismograms using hybrid techniques, combining deterministic long-period synthetics and shorter-period stochastic synthetics (e.g., Mai and Olsen 2006; Graves and Pitarka 2004; Liu et al. 2006). These methods are computationally relatively inexpensive and allow for the generation of synthetics with frequencies up to 10 Hz or higher. Using these techniques, a range of PGA estimates at the PBR locations can be found from an ensemble of realizations of broadband synthetics. Acknowledgements This project is funded by National Science Foundation (NSF) Information Technology Research through the award Division of Earth Sciences (EAR)-0122464 (The SCEC Community Modeling Environment: An Information Infrastructure for System-Level Earthquake Research). SCEC is funded by NSF Cooperative Agreement EAR- 0106924 and USGS Cooperative Agreement 02HQAG0008. This is SCEC publication #1143.

References Abrahamson NA, Somerville PG (1996) Effects of the hanging wall and footwall on ground motions recorded during the Northridge earthquake. Bull Seis Soc Am 86:593 599 Andrews JD (1999) Test of two methods for faulting in finitedifference calculations. Bull Seis Soc Am 89:931 937 Bell JW, Brune JN, Zeng Y (2004) Methodology for obtaining constraints of ground motion from precariously balanced rocks. Bull Seis Soc Am 94:285 303 Benites R, Olsen KB (2005) Modeling strong ground motion in the Wellington Metropolitan area, New Zealand. Bull Seis Soc Am 95:2180 2196 Day SM, Bradley CR (2001) Memory-efficient simulation of anelastic wave propagation. Bull Seism Soc Am 91:520 531 Freed AM, Lin J (2002) Accelerated stress buildup on the southern San Andreas Fault and surrounding regions caused by Mojave Desert earthquakes. Geology 30:571 574 Graves RW and Pitarka A (2004) Broadband time history simulation using a hybrid approach. Proceedings of the 13 th World Conference on Earthquake Engineering, Vancouver, Canada, August 1 6, 2004, Paper No. 1098. King G, Stein R, Lin J (1994) Static stress changes and the triggering of earthquakes. Bull Seis Soc Am 84:935 953 Liu P-C, Archuleta RJ, Hartzell SH (2006) Prediction of broadband ground-motion time histories: hybrid low/ high-frequency method with correlated random source parameters. Bull Seis Soc Am 96:2118 2130 Mai PM, Olsen KB (2006) Near-field broadband ground motions based on low-frequency finite-difference synthetics merged with high-frequency scattering operators. Seis Res Lett 77:300 Magistrale H, Day SM, Clayton R, Graves RW (2000) The SCEC southern California reference three-dimensional seismic velocity model version 2. Bull Seis Soc Am 90 (6B):S65 S76 Marcinkovich C, Olsen KB (2003) On the implementation of perfectly matched layers in a 3D fourth-order velocity-stress finite-difference scheme. J Geophys Res 108(B5):2276 Nicholson C (1996) Seismic behavior of the southern San Andreas Fault zone in the Coachella Valley, California: comparison of the 1948 and 1986 earthquake sequences. Bull Seis Soc Am 86:1331 1349 Oglesby DD, Dreger DS, Harris R, Ratchkovski N, Hansen R (2004) Inverse kinematic and forward dynamic models of the 2002 Denali, Alaska earthquake. Bull Seis Soc Am B6:S214 S233 Olsen KB (1994) Simulation of three-dimensional wave propagation in the Salt Lake Basin, Ph.D. Thesis, University of Utah, Salt Lake City, Utah, 157 pp Olsen KB, Day SM, Bradley CR (2003) Estimation of Q for long-period (>2 s) waves in the Los Angeles Basin. Bull Seis SocAm 93:627 638 Olsen KB, Day SM, Minster JB, Cui Y, Chourasia A, Faerman M, Moore R, Maechling P, Jordan T (2006) Strong shaking in Los Angeles expected from southern San Andreas earthquake. Geophys Res Lett 33:L07305 DOI 10.1029/2002JB002235 Olsen KB, Day SM, Minster JB, Cui Y, Chourasia A, Okaya D, Maechling P and Jordan T (2008) Spontaneous rupture simulations of Mw7.7 earthquakes on the southern San Andreas fault. Bull Seis Soc Am (in press) Peyrat S, Olsen KB, Madariaga R (2001) Dynamic modeling of the 1992 Landers earthquake. J Geophys Res 106:26,467 26,482 Purvance MD (2005) Overturning of slender blocks: numerical investigation and application to precariously balanced rocks in southern California, PhD Dissertation, University of Nevada, Reno Purvance MD, Brune JN, and Anooshehpoor A (2005) Attenuation relation consistency with precariously balanced rocks, Proceedings 2005 SCEC Conferenece, Palm Springs, CA, Sept. 11 14. Weldon R, Scharer K, Fumal T, Biasi G (2004) Wrightwood and the earthquake cycle: what a long recurrence record tells us about how faults work. Geol Seis Am 14:4 10 Working Group on California Earthquake Probabilities Seismic hazards in southern California: probable earthquakes, 1994 to 2024, 1995. Bull Seis Soc Am 85:379 439