Regional tectonic stress near the San Andreas fault in central and southern California

Similar documents
A mechanical model of the San Andreas fault and SAFOD Pilot Hole stress measurements

The San Andreas Fault Observatory at Depth: Recent Site Characterization Studies and the 2.2-Km-Deep Pilot Hole

Stress orientations at intermediate angles to the San Andreas Fault, California

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

Mechanics of Earthquakes and Faulting

Evidence for a strong San Andreas fault

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

Regional deformation and kinematics from GPS data

Does Aftershock Duration Scale With Mainshock Size?

Topographically driven groundwater flow and the San Andreas heat flow paradox revisited

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 113, B12415, doi: /2008jb005809, 2008

Velocity contrast along the Calaveras fault from analysis of fault zone head waves generated by repeating earthquakes

Predicted reversal and recovery of surface creep on the Hayward fault following the 1906 San Francisco earthquake

The generation of the first order stress field caused by plate tectonics

INFLUENCE OF LOCAL PERTURBATION ON REGIONAL STRESS AND ITS IMPACT ON THE DESIGN OF MAJOR UNDERGROUND STRUCTURE IN HYDROELECTRIC PROJECT

Source parameters II. Stress drop determination Energy balance Seismic energy and seismic efficiency The heat flow paradox Apparent stress drop

EARTHQUAKE LOCATIONS INDICATE PLATE BOUNDARIES EARTHQUAKE MECHANISMS SHOW MOTION

Mechanics of Earthquakes and Faulting

Geo736: Seismicity and California s Active Faults Introduction

Today: Basic regional framework. Western U.S. setting Eastern California Shear Zone (ECSZ) 1992 Landers EQ 1999 Hector Mine EQ Fault structure

Earth Science, (Tarbuck/Lutgens) Chapter 10: Mountain Building

SUPPLEMENTARY INFORMATION

Targeting of Potential Geothermal Resources in the Great Basin from Regional Relationships between Geodetic Strain and Geological Structures

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

Kinematics of the Southern California Fault System Constrained by GPS Measurements

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

Focused Observation of the San Andreas/Calaveras Fault intersection in the region of San Juan Bautista, California

Plate Boundary Observatory Working Group for the Central and Northern San Andreas Fault System PBO-WG-CNSA

Power-law distribution of fault slip-rates in southern California

Materials and Methods The deformation within the process zone of a propagating fault can be modeled using an elastic approximation.

GPS Strain & Earthquakes Unit 5: 2014 South Napa earthquake GPS strain analysis student exercise

Frictional behavior of materials in the 3D SAFOD volume

Measurements in the Creeping Section of the Central San Andreas Fault

EXAMINATION ON CONSECUTIVE RUPTURING OF TWO CLOSE FAULTS BY DYNAMIC SIMULATION

by George E. Hilley, J Ramón Arrowsmith, and Elizabeth Stone Introduction

Part 2 - Engineering Characterization of Earthquakes and Seismic Hazard. Earthquake Environment

Summary so far. Geological structures Earthquakes and their mechanisms Continuous versus block-like behavior Link with dynamics?

Apparent and True Dip

Aftershock Sequences Modeled with 3-D Stress Heterogeneity and Rate-State Seismicity Equations: Implications for Crustal Stress Estimation

Jack Loveless Department of Geosciences Smith College

Comment on A new estimate for present-day Cocos-Caribbean plate motion: Implications

High-resolution image of Calaveras Fault seismicity

of other regional earthquakes (e.g. Zoback and Zoback, 1980). I also want to find out

Mechanics of faulting

Stress Orientations Obtained from Earthquake Focal Mechanisms: What Are Appropriate Uncertainty Estimates?

Earthquake and Volcano Deformation

Geophysical Journal International

Brittle Deformation. Earth Structure (2 nd Edition), 2004 W.W. Norton & Co, New York Slide show by Ben van der Pluijm

UCERF3 Task R2- Evaluate Magnitude-Scaling Relationships and Depth of Rupture: Proposed Solutions

Seismic and aseismic processes in elastodynamic simulations of spontaneous fault slip

Force and Stress. Processes in Structural Geology & Tectonics. Ben van der Pluijm. WW Norton+Authors, unless noted otherwise 1/9/ :35 PM

Structural deformation across the southwest Mina deflection, California-Nevada: Field studies in the Huntoon Springs area.

The Mechanics of Earthquakes and Faulting

Fault Processes on the Anza section of the San Jacinto Fault

The Frictional Regime

Weakness of the Intracontinental Strike-Slip Kunlun Fault and Implications for Tibetan Tectonics

Verification of the asperity model using seismogenic fault materials Abstract

Overview of Seismic Source Characterization for the Diablo Canyon Power Plant

Lack of continuity of the San Andreas Fault in southern California: Three-dimensional fault models and earthquake scenarios

GPS Strain & Earthquakes Unit 4: GPS strain analysis examples Student exercise

Deformation of Rocks. Orientation of Deformed Rocks

In geophysics there is increasing interest in modeling spatially nonuniform stress,

CONTENTS PREFACE. VII 1. INTRODUCTION VARIOUS TOPICS IN SEISMOLOGY TECTONICS PERTAINING TO EQ PREDICTION 5

Elastic Rebound Theory

Initiation of the San Jacinto Fault and its Interaction with the San Andreas Fault: Insights from Geodynamic Modeling

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

Stress modulation on the San Andreas fault by interseismic fault system interactions Jack Loveless and Brendan Meade, Geology, 2011

Numerical modeling of strike-slip creeping faults and implications for the Hayward fault, California

DEFORMATION KINEMATICS OF TIBETAN PLATEAU DETERMINED FROM GPS OBSERVATIONS

Evolution of stress in Southern California for the past 200 years from coseismic, postseismic and interseismic stress changes

Ground displacement in a fault zone in the presence of asperities

State of Stress from Focal Mechanisms Before and After the 1992 Landers Earthquake Sequence

External Grant Award Number 04HQGR0058 IMPROVED THREE-DIMENSIONAL VELOCITY MODELS AND EARTHQUAKE LOCATIONS FOR CALIFORNIA

by Deborah Elaine Smith * and Thomas H. Heaton Introduction

A Model of Three Faults

BEM Model of slip on the Channel Islands Thrust, CA

Lecture 2: Deformation in the crust and the mantle. Read KK&V chapter 2.10

The Pennsylvania State University. The Graduate School. College of Earth and Mineral Sciences

San Andreas Movie Can It Happen?

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

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

Tectonic stress field in rift systems a comparison of Rhinegraben, Baikal Rift and East African Rift

Lateral extrusion and tectonic escape in Ilan Plain of northeastern Taiwan

State of Stress in Seismic Gaps Along the SanJacinto Fault

Estimating fault slip rates, locking distribution, elastic/viscous properites of lithosphere/asthenosphere. Kaj M. Johnson Indiana University

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

Lecture # 6. Geological Structures

Quaternary Processes of the Monterey Bay Area Field Trip Notes

Earthquakes in Barcelonnette!

Overview of the Seismic Source Characterization for the Palo Verde Nuclear Generating Station

Chapter 15 Structures

Utah State University. Ryan M. Lee Utah State University

Surface changes caused by erosion and sedimentation were treated by solving: (2)

Material is perfectly elastic until it undergoes brittle fracture when applied stress reaches σ f

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

Current stress state and principal stress rotations in the vicinity of the Chelungpu fault induced by the 1999 Chi-Chi, Taiwan, earthquake

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

Possibility of reservoir induced seismicity around three gorges dam on Yangtze river

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

Activity Pacific Northwest Tectonic Block Model

Transcription:

GEOPHYSICAL RESEARCH LETTERS, VOL. 31, L15S11, doi:10.1029/2003gl018918, 2004 Regional tectonic stress near the San Andreas fault in central and southern California J. Townend School of Earth Sciences, Victoria University of Wellington, Wellington, New Zealand M. D. Zoback Department of Geophysics, Stanford University, Stanford, California, USA Received 23 October 2003; revised 18 December 2003; accepted 30 December 2003; published 29 July 2004. [1] Throughout central and southern California, a uniform NNE-SSW direction of maximum horizontal compressive stress is observed that is remarkably consistent with the superposition of stresses arising from lateral variations in lithospheric buoyancy in the western United States, and farfield Pacific-North America plate interaction. In central California, the axis of maximum horizontal compressive stress lies at a high angle to the San Andreas fault (SAF). Despite relatively few observations near (±10 km) the fault, observations in the greater San Francisco Bay area indicate an angle of as much as 85, implying extremely low fault strength. In southern California, observations of stress orientations near the SAF are rotated slightly counterclockwise with respect to the regional field. Nevertheless, we observe an approximately constant angle between the SAF and the maximum horizontal stress direction of 68 ± 7 along 400 km of the fault, indicating that the SAF has moderately low frictional strength in southern California. INDEX TERMS: 8010 Structural Geology: Fractures and faults; 8020 Structural Geology: Mechanics; 8123 Tectonophysics: Dynamics, seismotectonics; 8164 Tectonophysics: Stresses crust and lithosphere. Citation: Townend, J., and M. D. Zoback (2004), Regional tectonic stress near the San Andreas fault in central and southern California, Geophys. Res. Lett., 31, L15S11, doi:10.1029/2003gl018918. 1. Introduction [2] It has been known for some time that stress orientations along the San Andreas fault (SAF) [Mount and Suppe, 1987; Zoback et al., 1987; Jones, 1988; Townend and Zoback, 2001] and the lack of a distinct heat flow anomaly at the trace of the fault [Lachenbruch and Sass, 1992, and references therein; Saffer et al., 2003] indicate average shear tractions less than 20 25 MPa in the seismogenic upper crust. These observations are difficult to reconcile with typical frictional coefficients of common rock types (m =0.6 1.0) such as are measured in laboratory settings [Byerlee, 1978] or inferred from deep borehole stress measurements [Townend and Zoback, 2000], which suggest shear tractions approximately five times as large. High frictional coefficients (commonly referred to as Byerlee friction ) imply an angle (b) between the axis of maximum Copyright 2004 by the American Geophysical Union. 0094-8276/04/2003GL018918$05.00 horizontal compression, S Hmax, and the strike of a vertical strike-slip fault such as the San Andreas of 30 35. Instead, what is typically observed throughout California is a much larger angle indicating near fault-normal compression in some places; such observations are especially noteworthy when the data come from crustal volumes immediately adjacent to the SAF [Jones, 1988; Oppenheimer et al., 1988; Zoback and Beroza, 1993; Townend and Zoback, 2001]. [3] The observed long-range (1000 5000 km) uniformity of stress orientations and relative magnitudes in intraplate regions suggests that plate-driving forces provide the largest component of the total stress field [Zoback et al., 1989; Zoback, 1992]. However, in some areas, including the western United States, stresses caused by lateral variations in crustal buoyancy appear to provide a large component of the horizontal stress field [Fleitout and Froidevaux, 1982; Jones et al., 1996]. How large, plate-bounding faults such as the SAF accomplish the slip required of them by global kinematics given ambient stresses is an outstanding question in crustal mechanics. In this study we compare different indicators of tectonic stress directions in central and southern California in order to put planned in situ stress measurements from the San Andreas Fault Observatory at Depth (SAFOD) drilling project in a regional context. 2. Central California [4] Figure 1 illustrates crustal stress orientations in central California obtained from focal mechanism inversions [Townend and Zoback, 2001] and a variety of sources represented in the World Stress Map Database, most notably stress-induced wellbore breakouts in oil and gas wells [Fuchs and Müller, 2001]. The dashed trajectories are interpolations of S Hmax directions calculated by Flesch et al. [2000] using a variational technique to estimate the minimum deviatoric horizontal stress field resulting from lateral variations in gravitational potential energy (or lithospheric buoyancy) and deformation due to interaction between the North American and Pacific plates. Borehole breakout data, hydraulic fracture data, and focal mechanism stress inversion results reveal highly consistent S Hmax orientations at adjacent locations, a strong correlation with the computed S Hmax trajectories, and a very high angle between the S Hmax direction and the SAF (b 85 ) along the San Francisco Peninsula. It is noteworthy that the Hayward, Calaveras and San Gregorio faults each appear to be as L15S11 1of5

Figure 1. Maximum horizontal compression (S Hmax ) and crustal velocity data from central California shown in an oblique Mercator projection about the North America-Pacific Euler pole [DeMets et al., 1990]. S Hmax directions determined from borehole breakouts are shown by inward-pointing arrows, from hydraulic fracturing experiments by stars, and from earthquake focal mechanism inversions by split circles [Fuchs and Müller, 2001; Townend and Zoback, 2001]. The dashed trajectories show regional S Hmax directions calculated using a model of lithospheric buoyancy and plate interaction [Flesch et al., 2000], and the vectors illustrate crustal velocity data relative to North America [Murray and Segall, 2001]. The numbers in parentheses are Quaternary fault slip rates [Working Group on Earthquake Probabilities, 1999]. The SAFOD drillhole location is marked by a white circle. SF San Francisco. poorly oriented with respect to the contemporary stress field as the SAF itself [Schaff et al., 2002; Provost and Houston, 2003]. [5] The very high value of b 85 observed along the San Francisco Peninsula is particularly interesting because the stress data were obtained from boreholes and focal mechanism stress inversions as close as 1 km to the fault s surface trace [Townend and Zoback, 2001]. Consequently, the ratio of shear to normal traction acting on the SAF in this area is very low. Assuming that the crust adjacent to the SAF is critically stressed [Townend and Zoback, 2000], we estimate the average shear traction acting on planes parallel to the SAF (evaluated at a depth of 7.5 km in a strike-slip stress regime using a crustal friction coefficient of 0.6) to be as little as 10 MPa along the San Francisco Peninsula (b = 85 ), the effective normal traction to be 180 MPa, and the coefficient of friction to be an extremely low 0.06. These estimates are compatible with those of Mount and Suppe [1987] and Zoback et al. [1987]. [6] Near Parkfield, the orientation of the regional stress field is highly compatible with stress measurements made in the SAFOD Pilot Hole [Hickman and Zoback, 2004], which indicate a clockwise rotation of S Hmax with depth to an angle of b =70 at 2.0 2.2 km. 3. Southern California [7] A map of southern California stress orientations is shown in Figure 2. As in central California, several important features of this map deserve specific mention: first, individual data sets produce remarkably consistent S Hmax estimates at relatively fine scales, as demonstrated particularly well by the cluster of near-parallel stress orientations obtained by focal mechanism stress inversion of seismicity that occurred after the 1994 Northridge earthquake (241.4 E, 34.3 N), in the southern Sierra Nevada (242.2 E, 35.9 N) and eastern Mojave Desert (243.0 E, 35.0 N), and by the clusters of borehole breakout and hydrofracture results in the southern San Joaquin Valley. Second, the results obtained using different techniques generally agree very well, in southern California as a whole but particularly in the southern San Joaquin Valley and near Los Angeles (241.6 E, 34.0 N). Third, S Hmax exhibits a relatively uniform NNE-SSW orientation at regional scales that is at approximately 60 70 to the average strike of the SAF throughout southern California. Finally, there is remarkably good agreement between the observed stress field and the dynamical modeling results obtained independently by Flesch et al. [2000]. [8] No stress rotations near the SAF are apparent at the scale of these observations. An independent study that incorporates focal mechanism uncertainties explicitly [Abers and Gephart, 2001] substantiates this. The rose diagram inset in Figure 2 summarizes estimates of b at 70 locations within 10 km of the SAF surface trace: the mean angle between S Hmax and the local fault trace is 68 ± 7. S Hmax is oriented at a relatively consistent angle to the SAF throughout southern California, a result also noted on the basis of a smaller data set by Jones [1988], suggesting that the stress field and fault geometry have evolved in a 2of5

Figure 2. Orientation of the axis of maximum horizontal compression (S Hmax ) in southern California. Plain lines indicate S Hmax data from earthquake focal mechanism inversions; all other symbols are the same as in Figure 1. The inset summarizes the angle between S Hmax and the local fault strike at points within 10 km of the SAF. EMSZ Eastern Mojave Shear Zone. self-organized manner. The corresponding orientation is essentially the same (63 ± 10 ) if only those data within 5 km of the fault trace are considered (34 data points). Such high b angles are not compatible with Andersonian faulting theory based on Byerlee friction and hydrostatic fluid pressures; rather, they support the contention that the SAF is subject to a significant component of fault-normal compression and therefore slips at low shear/normal stress ratios. The average shear and effective normal tractions on planes parallel to the SAF in southern California are estimated to be 50 MPa and 150 MPa, and 30 MPa and 170 MPa, for angles of 61 and 75, respectively: these correspond to effective friction coefficients of 0.2 0.3. 4. Discussion and Summary [9] The stress data reveal that strike-slip displacement along the SAF and subsidiary structures occurs at an angle of approximately 68 to the axis of maximum horizontal compression in southern California, and as high as 85 in central California. In situ stress data and dynamical modeling results from southern and central California exhibit very good agreement in terms of principal tectonic stress directions. [10] In modeling the contributions of lithospheric buoyancy and internal plate deformation to the horizontal deviatoric stress field, Flesch et al. [2000] made no attempt to match in situ stress data, but the agreement between their computational results and the observed stress directions at regional scales is very good throughout most of California. At distances greater than 10 km from the SAF fault trace, the average discrepancy between the S Hmax directions determined from borehole or earthquake data and Flesch et al. s results is 8, with the latter being rotated slightly clockwise with respect to the observed stress directions. Nevertheless, it should be noted that there are several areas where the discrepancy is particularly marked. The majority of the locations with high discrepancies are located in one of three distinct clusters: the Big Bend region near (241.0 E, 34.8 N), north of the Garlock fault at (241.5 E, 35.4 N), and near San Bernadino at (242.3 E, 34.2 N). The Big Bend and San Bernadino knots both coincide with appreciable, localized bends in the strike of the SAF s surface trace; the San Bernadino cluster also corresponds to the southernmost rupture of the 1857 Fort Tejon earthquake [Jones, 1988]. The Garlock cluster is at the east end of the source region of the 1952 Kern County earthquake (M7.8), which has been previously interpreted to have substantially affected the local stress field [Castillo and Zoback, 1995]. [11] The fact that the angle between S Hmax and the SAF remains approximately constant over a 400 km distance in 3of5

southern California, despite pronounced changes in the fault s strike, indicates a degree of self-organization or interaction between the tectonic stress field and the fault. If this were not the case, then the relatively uniform orientation of the regional stress field would result in b varying systematically with the fault s strike. By comparing discrepancies between the observed and modeled stress orientations close to (10 km) and far away from the SAF, we have estimated the effect of this interaction to be equivalent to an anticlockwise rotation of S Hmax by less than 7. [12] In southern California, specifically along the section of the SAF that ruptured in the M8+ 1857 Fort Tejon earthquake, Hardebeck and Hauksson [1999] suggested that S Hmax rotates by 40 within ±20 km of the SAF to an angle at the fault trace of b =45. This apparent rotation has been used in support of a model positing the SAF s average frictional strength to be consistent with Byerlee friction and hydrostatic fluid pressures [Scholz, 2000], and hence to be comparable to the strength of the adjacent crust. Hardebeck and Hauksson [1999, 2001] suggested to the contrary that while the SAF and adjacent crust have similar strengths, these must necessarily be much lower than predicted using a critically stressed crust model and Byerlee friction. [13] We are of the opinion that any rotation of S Hmax to such low angles must occur over a horizontal distance less than our observation scale of 5 km. Indeed, the similarity between focal mechanism-based estimates of S Hmax near Parkfield and the deepest SAFOD Pilot Hole measurements, which were made only 1.8 km from the SAF, suggests that any rotation must take place on even smaller horizontal scales. It is important to emphasize that while we cannot rule out average stress orientations at the SAF fault plane itself of b < 68, existing focal mechanism inversion results and borehole data do not substantiate that interpretation. Hardebeck and Hauksson [2001] demonstrated that the focal mechanism stress inversion results they had previously obtained [Hardebeck and Hauksson, 1999] were very similar to those of Townend and Zoback [2001], in which case we conclude that disagreement over the near-field value of b stems from the comparison of a small number of results from separate cross-sections with a larger sample of 70 samples collected over a more representative length (400 km) of the SAF. [14] Overall, it is striking that the observed state of stress in southern and central California is so consistent over such a large area, although both the borehole stress measurements and inversions of independent sets of earthquake focal mechanisms reveal gradual lateral stress variations throughout the region. In general, the observed stress field is very similar to Flesch et al. s [2000] calculations of buoyancy-derived stresses associated with the thermally uplifted Basin and Range province and far-field shear related to Pacific-North America relative plate motion, and consistent with their findings based on regional variations in effective viscosity that the SAF is a mechanically weak boundary. Thus, while GPS data show that motions of the crustal blocks adjacent to the plate boundary are consistent with local plate kinematics, the forces acting within the crust seem to be controlled by relatively far-field processes. The consistently high angle between S Hmax and the SAF observed along a 400 km-long section of the fault in southern California (and the even higher angles seen in the sparser data set further north) demonstrates that the SAF s frictional resistance to slip though low in both absolute and relative senses is nevertheless manifest in the tectonic field immediately adjacent to the fault. [15] Acknowledgments. We are grateful to Lucy Flesch and Bill Holt for making their results available, and to Naomi Boness, Björn Lund, and George Thompson for comments. GMT version 3.3.6 was used to construct both figures. References Abers, G. A., and J. W. Gephart (2001), Direct inversion of earthquake first motions for both the stress tensor and focal mechanisms and application to southern California, J. Geophys. Res., 106(B11), 26,523 26,540. Byerlee, J. (1978), Friction of rocks, Pure Appl. Geophys., 116, 615 626. Castillo, D. A., and M. D. Zoback (1995), Systematic stress variations in the southern San Joaquin Valley and along the White Wolf fault: Implications for the rupture mechanics of the 1952 M S 7.8 Kern County earthquake and contemporary seismicity, J. Geophys. Res., 100(B4), 6249 6264. DeMets, C., R. G. Gordon, D. F. Argus, and S. Stein (1990), Current plate motions, Geophys. J. Int., 101, 425 478. Fleitout, L., and C. Froidevaux (1982), Tectonics and topography for a lithosphere containing density heterogeneities, Tectonics, 1, 21 56. Flesch, L. M., W. E. Holt, A. J. Haines, and B. Shen-Tu (2000), Dynamics of the Pacific-North American plate boundary in the western United States, Science, 287, 834 836. Fuchs, K., and B. Müller (2001), World stress map of the Earth: A key to tectonic processes and technological applications, Naturwissenschaften, 88, 357 371. Hardebeck, J. L., and E. Hauksson (1999), Role of fluids in faulting inferred from stress field signatures, Science, 285, 236 239. Hardebeck, J. L., and E. Hauksson (2001), Crustal stress field in southern California and its implications for fault mechanics, J. Geophys. Res., 106(B10), 21,859 21,882. Hickman, S., and M. D. Zoback (2004), Stress orientations and magnitudes in the SAFOD Pilot Hole, Geophys. Res. Lett., 31, L15S12, doi:10.1029/ 2004GL020043. Jones, L. M. (1988), Focal mechanisms and the state of stress on the San Andreas fault in southern California, J. Geophys. Res., 93, 8869 8891. Jones, C. H., J. R. Unruh, and L. J. Sonder (1996), The role of gravitational potential energy in active deformation in the southwestern United States, Nature, 381, 37 41. Lachenbruch, A. H., and J. H. Sass (1992), Heat flow from Cajon Pass, fault strength, and tectonic implications, J. Geophys. Res., 97(B4), 4995 5015. Mount, V. S., and J. Suppe (1987), State of stress near the San Andreas fault: Implications for wrench tectonics, Geology, 15, 1143 1146. Murray, M. H., and P. Segall (2001), Modeling broadscale deformation in northern California and Nevada from plate motions and elastic strain accumulation, Geophys. Res. Lett., 28(22), 4315 4318. Oppenheimer, D. H., P. A. Reasenberg, and R. W. Simpson (1988), Fault plane solutions for the 1984 Morgan Hill, California, earthquake sequence: Evidence for the state of stress on the Calaveras fault, J. Geophys. Res., 93(B8), 9007 9026. Provost, A.-S., and H. Houston (2003), Stress orientations in northern and central California: Evidence for the evolution of frictional strength along the San Andreas plate boundary system, J. Geophys. Res., 108(B3), 2175, doi:10.1029/2001jb001123. Saffer, D. M., B. A. Bekins, and S. Hickman (2003), Topographically driven groundwater flow and the San Andreas heat flow paradox revisited, J. Geophys. Res., 108(B5), 2274, doi:10.1029/2002jb001849. Schaff, D. P., G. H. R. Bokelmann, G. C. Beroza, F. Waldhauser, and W. L. Ellsworth (2002), High-resolution image of Calaveras fault seismicity, J. Geophys. Res., 107(B9), 2186, doi:10.1029/2001jb000633. Scholz, C. H. (2000), Evidence for a strong San Andreas fault, Geology, 28, 163 166. Townend, J., and M. D. Zoback (2000), How faulting keeps the crust strong, Geology, 28, 399 402. Townend, J., and M. D. Zoback (2001), Implications of earthquake focal mechanisms for the frictional strength of the San Andreas fault system, in The Nature and Significance of Fault Zone Weakening, edited by R. E. Holdsworth et al., Geol. Soc. London Spec. Publ., 186, 13 21. Working Group on Earthquake Probabilities (1999), Earthquake probabilities in the San Francisco Bay region: 2000 2030 A summary of findings, U. S. Geol. Surv. Open File Rep., 99-517, 60 pp. 4of5

Zoback, M. D., and G. C. Beroza (1993), Evidence for near-frictionless faulting in the 1989 (M6.9) Loma Prieta, California, earthquake and its aftershocks, Geology, 21, 181 185. Zoback, M. L. (1992), First- and second-order patterns of stress in the lithosphere: The World Stress Map Project, J. Geophys. Res., 97(B8), 11,703 11,728. Zoback, M. D., et al. (1987), New evidence for the state of stress on the San Andreas fault system, Science, 238, 1105 1111. Zoback, M. L., et al. (1989), Global patterns of tectonic stress, Nature, 341, 291 298. J. Townend, School of Earth Sciences, Victoria University of Wellington, P.O. Box 600, Wellington, New Zealand. (john.townend@vuw.ac.nz) M. D. Zoback, Department of Geophysics, Stanford University, Stanford, CA 94305-2215, USA. (zoback@pangea.stanford.edu) 5of5