Which Dynamic Rupture Parameters Can Be Estimated from Strong Ground Motion and Geodetic Data?

Similar documents
Distinguishing barriers and asperities in near-source ground motion

Criticality of Rupture Dynamics in 3-D

Simulation of earthquake rupture process and strong ground motion

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

Improvement in the Fault Boundary Conditions for a Staggered Grid Finite-difference Method

The Length to Which an Earthquake will go to Rupture. University of Nevada, Reno 89557

Spectral Element simulation of rupture dynamics

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

Title from Slip-Velocity Functions on Ear.

Effect of normal stress during rupture propagation along nonplanar faults

Ground displacement in a fault zone in the presence of asperities

Dynamic source inversion for physical parameters controlling the 2017 Lesvos earthquake

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

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

Effective friction law for small scale fault heterogeneity in 3D dynamic rupture

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

The role of fault continuity at depth in numerical simulations of earthquake rupture

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

Earthquake stress drop estimates: What are they telling us?

Slip-weakening behavior during the propagation of dynamic ruptures obeying rate- and state-dependent friction laws

Dynamic modeling of the 1992 Landers earthquake

Earthquake Stress Drops in Southern California

Fault Representation Methods for Spontaneous Dynamic Rupture Simulation. Luis A. Dalguer

Fault Specific, Dynamic Rupture Scenarios for Strong Ground Motion Prediction

(Somerville, et al., 1999) 2 (, 2001) Das and Kostrov (1986) (2002) Das and Kostrov (1986) (Fukushima and Tanaka, 1990) (, 1999) (2002) ( ) (1995

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

The combined inversion of seismic and geodetic data for the source process of the 16 October,

Synthetic Seismicity Models of Multiple Interacting Faults

Fracture surface energy of natural earthquakes from the viewpoint of seismic observations

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

To Editor of Bulletin of the Seismological Society of America

Heterogeneous distribution of the dynamic source parameters of the 1999 Chi-Chi, Taiwan, earthquake

Numerical study on multi-scaling earthquake rupture

Seismic and aseismic processes in elastodynamic simulations of spontaneous fault slip

Regional Geodesy. Shimon Wdowinski. MARGINS-RCL Workshop Lithospheric Rupture in the Gulf of California Salton Trough Region. University of Miami

Postseismic Deformation from ERS Interferometry

The Combined Inversion of Seismic and Geodetic Data for the Source Process of the 16 October 1999 M w 7.1 Hector Mine, California, Earthquake

Source Process and Constitutive Relations of the 2011 Tohoku Earthquake Inferred from Near-Field Strong-Motion Data

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

EFFECTS OF NON-LINEAR WEAKENING ON EARTHQUAKE SOURCE SCALINGS

Constraining fault constitutive behavior with slip and stress heterogeneity

EXAMINATION ON CONSECUTIVE RUPTURING OF TWO CLOSE FAULTS BY DYNAMIC SIMULATION

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

EFFECTS OF CORRELATION OF SOURCE PARAMETERS ON GROUND MOTION ESTIMATES

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

Properties of Dynamic Earthquake Ruptures With Heterogeneous Stress Drop

Constraining earthquake source inversions with GPS data: 2. A two-step approach to combine seismic and geodetic data sets

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

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

Spatial and Temporal Distribution of Slip for the 1999 Chi-Chi, Taiwan, Earthquake

High Resolution Imaging of Fault Zone Properties

ABSTRACT. Key words: InSAR; GPS; northern Chile; subduction zone.

Effects of fault geometry and slip style on near-fault static displacements caused by the 1999 Chi-Chi, Taiwan earthquake

Basics of the modelling of the ground deformations produced by an earthquake. EO Summer School 2014 Frascati August 13 Pierre Briole

Dynamic source modeling of the 1978 and 2005 Miyagi oki earthquakes: Interpretation of fracture energy

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

Challenges in earthquake physics and source imaging

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

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

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

Imaging Earthquake Source Complexity

Dynamic stress field of a kinematic earthquake source model

Conditions governing the occurrence of supershear ruptures under slip-weakening friction

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

Homogeneous vs. realistic heterogeneous material-properties in subduction zone models: Coseismic and postseismic deformation

Calculation of the local rupture speed of dynamically propagating earthquakes

Depth variation of coseismic stress drop explains bimodal earthquake magnitude-frequency distribution

NOTES AND CORRESPONDENCE Segmented Faulting Process of Chelungpu Thrust: Implication of SAR Interferograms

BROADBAND SIMULATION FOR A HYPOTHETICAL M W 7.1 EARTHQUAKE ON THE ENRIQUILLO FAULT IN HAITI

Development of a Predictive Simulation System for Crustal Activities in and around Japan - II

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

FRICTIONAL HEATING DURING AN EARTHQUAKE. Kyle Withers Qian Yao

Surface Rupture in Kinematic Ruptures Models for Hayward Fault Scenario Earthquakes

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

Nonlinear site response from the 2003 and 2005 Miyagi-Oki earthquakes

Inversion of Earthquake Rupture Process:Theory and Applications

Slip distributions of the 1944 Tonankai and 1946 Nankai earthquakes including the horizontal movement effect on tsunami generation

Coseismic Deformation from the 1999 M w 7.1 Hector Mine, California, Earthquake as Inferred from InSAR and GPS Observations

Mechanics of Earthquakes and Faulting

Investigation of the Rupture Process of the 28 June 1992 Landers Earthquake Utilizing TERRAscope

Evidence for a Supershear Transient during the 2002 Denali Fault Earthquake

Preliminary slip model of M9 Tohoku earthquake from strongmotion stations in Japan - an extreme application of ISOLA code.

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

Supporting Information for Inferring field-scale properties of a fractured aquifer from ground surface deformation during a well test

Numerical Modeling of Earthquake Dynamic Rupture: Requirements for Realistic Modeling

Nonuniform prestress from prior earthquakes and the effect on dynamics of branched fault systems

Source Description of the 1999 Hector Mine, California, Earthquake, Part I: Wavelet Domain Inversion Theory and Resolution Analysis

MODELING OF HIGH-FREQUENCY WAVE RADIATION PROCESS ON THE FAULT PLANE FROM THE ENVELOPE FITTING OF ACCELERATION RECORDS

Originally published as:

Numerical simulation of seismic cycles at a subduction zone with a laboratory-derived friction law

Dynamic stress field of a kinematic earthquake source model with k-squared slip distribution

Pulse like, crack like, and supershear earthquake ruptures with shear strain localization

Hitoshi Hirose (1), and Kazuro Hirahara (2) Abstract. Introduction

Earthquakes and Faulting

4. Numerical algorithms for earthquake rupture dynamic modeling

Characterization of nucleation during laboratory earthquakes

AN INTRODUCTION TO EARTHQUAKE GEODESY : Another Effort for Earthquake Hazard Monitoring

CHARACTERIZING EARTHQUAKE SLIP MODELS FOR THE PREDICTION OF STRONG GROUND MOTION

Contribution of HPC to the mitigation of natural risks. B. Feignier. CEA-DAM Ile de France Département Analyse, Surveillance, Environnement

2 compared with the uncertainty in relative positioning of the initial triangulation (30 { 50 mm, equivalent to an angular uncertainty of 2 p.p.m. ove

Transcription:

Pure appl. geophys. 161 (24) 2155 2169 33 4553/4/122155 15 DOI 1.17/s24-4-2555-9 Ó Birkhäuser Verlag, Basel, 24 Pure and Applied Geophysics Which Dynamic Rupture Parameters Can Be Estimated from Strong Ground Motion and Geodetic Data? SOPHIE PEYRAT 1,KIM B. OLSEN 1, and RAU L MADARIAGA 2 Abstract We have tested to which extent commonly used dynamic rupture parameters can be resolved for a realistic earthquake scenario from all available observations. For this purpose we have generated three dynamic models of the Landers earthquake using a single, vertical, planar fault with heterogeneity in either the initial stress, the yield stress, or the slip-weakening distance. Although the dynamic parameters for these models are inherently different, all the simulations are in agreement with strong motion, GPS, InSAR, and field data for the event. The rupture propagation and slip distributions obtained for each model are similar, showing that the solution of the dynamic problem is non-unique. In other words, it is not always possible to separate strength drop and the slip-weakening distance using rupture modeling, in agreement with the conclusions by GUATTERI and SPUDICH (2). Key words: Rupture dynamics, 1992 Landers earthquake, initial stress, friction law parameters, strong motion data, geodetic data. Introduction Earthquake rupture is initiated when the stress on a pre-existing fault reaches a critical level on a finite patch of the fault. After initiation the constitutive laws and the initial stress on the fault control how the earthquake propagates and arrests on the fault. Thus the initial stress (T e ) and the parameters of the constitutive laws (the yield stress T u and the slip-weakening distance D c in our case) are usually used as the initial parameters in dynamic rupture models. In addition, rupture propagation is characterized by various other parameters, such as slip (in part observable in case of surface rupture) and slip velocity which is responsible for the generation of seismic waves as well as ground failure caused by the earthquake. The level and variation of the initial stress and friction affects the slip, slip velocity and stress drop in a strongly nonlinear fashion, as demonstrated by simulations of dynamic rupture on heterogeneous faults (i.e., OLSEN et al., 1997; PEYRAT et al., 21). For example, larger values of the frictional parameters generate 1 Institute for Crustal Studies, University of California, Santa Barbara, Santa Barbara, CA 9316-11, U.S.A. E-mail: peyrat@crustal.ucsb.edu, kbolsen@crustal.ucsb.edu 2 Laboratoire de Géologie, Ecole Normale Supérieure, 24 rue Lhomond, 75231 Paris Cedex 5, France. E-mail: madariag@geologie.ens.fr

2156 Sophie Peyrat et al. Pure appl. geophys., larger fracture energy (G /ðt u T f ÞD c, where T f is the kinematic friction) and therefore increase rupture resistance. On the other hand, larger values of the initial stress generate more available strain energy ðu / Te 2 Þ and promote rupture. In addition, for realistic scenarios where the initial conditions are heterogeneous the behavior of the rupture is rather difficult to estimate. The dynamic rupture and therefore the radiated waves are not only controlled by the values of the initial conditions, but also by the associated characteristic length scales. Two classical and idealized models describe heterogeneous fault rupture. DAS and AKI (1977) proposed the barrier model, where the fault contains areas of increased rupture resistance. The nucleation occurs in a region of weak rupture resistance and the rupture propagates between barriers. Alternatively, KANAMORI and STEWART (1978) proposed the asperity model, where the fault contains areas of stress concentrations due to former earthquakes or aseismic slip. The barrier and asperity models are viewed as complementary in describing dynamic rupture propagation. GUATTERI and SPUDICH (2) demonstrated with a quasi-dynamical study that it is not possible to estimate strength drop and the slip weakening distance separately from low-frequency strong motion data (<1.6 Hz) only for a realistic earthquake through kinematic modeling. In our study we take a step further to examine to which extent the parameters can be estimated from trial-and-error inversion of dynamic rupture, in both cases of asperity and barrier models. The goal of this study is to understand the fundamental role of the different dynamic parameters involved during earthquake rupture, as well as their signature in seismic, geodetic and field data. Do barrier and asperity models generate differences in accelerograms, slip vectors, fringes in interferometric data and/or in surface slip? Which (if any) rupture parameters can we constrain from these types of observed data? In order to answer these questions, we use dynamic modeling of the M 7.3 June 28, 1992, Landers earthquake (Fig. 1), one of the largest, best-recorded strike-slip earthquakes in California. In a first attempt at constructing a dynamic model of the Landers earthquake, OLSEN et al. (1997) studied the frictional conditions under which rupture could propagate and then modeled the dynamic rupture process for the initial stress field obtained from the slip distribution determined by WALD and HEATON (1994). PEYRAT et al. (21) carried the initial study by OLSEN et al. (1997) a step further by inverting for the initial conditions on the fault using the observed ground motion. By trial-and-error inversion of strong motion data PEYRAT et al. (21, 22) constructed an asperity model (variable initial stress T e ) and a barrier model (variable yield stress T u ) for the Landers rupture. However, there is evidence from inversion of seismic data for large earthquakes (i.e., IDE and TAKEO, 1997, for the 1995 Kobe earthquake), that the slip-weakening distance D c may vary spatially. Therefore, we build an additional barrier model, with variable slip-weakening distance D c, from trial-and-error inversion of strong motion data. The barrier and asperity models were related using the non-dimensional parameter j identified by MADARIAGA and OLSEN (2) (see also DUNHAM et al., 23), which contains information about all the

Vol. 161, 24 Which Dynamic Rupture Parameters can be Estimated 2157-118 -117-116 -115 FTI BKR 35 35 BRN BRS YER 34 34 PSA H5 INI 33-118 -117-116 33-115 km 4 8 Figure 1 Map of the Landers earthquake region with the fault trace and the location of the strong motion stations (triangles) used in this study. initial parameters studied here (T e, T u and D c ). GUATTERI and SPUDICH (2) used the fracture energy G for this purpose, which however is insufficient for our dynamic models. Finally, in order to validate our models and examine the signature of these models in different measurable data sets, we compare the synthetic Global Positioning System (GPS), interferometric synthetic aperture radar (InSAR), and surface slip data for the three models to the measured data. Numerical Modeling Method and Friction Law We invert the seismic data for the Landers earthquake using a well-posed dynamic fault model, consisting of spontaneous rupture propagation along a vertical, planar fault. We solve the elastodynamic equation using a fourth-order staggered-grid finite difference method in a three-dimensional medium (MADARIAGA et al., 1998). The elastodynamic equations are combined with a free surface boundary condition at the top of the fault and absorbing boundary conditions introduced at the grid edges to eliminate artificial reflections. Finally, the fault is modeled as an internal boundary on which stress is related to slip by a friction law. Here we use the simple slip-weakening law ( T ðdþ ¼ ðt u T f Þ 1 D D c þ T f D D c ; ð1þ T f D > D c where T u is the yield stress, T f is the kinematic friction at high slip and D c is the slipweakening distance. This friction law was introduced by IDA (1972). Slip is zero until the stress reaches a peak value (T u ); then it starts to increase while, simultaneously,

2158 Sophie Peyrat et al. Pure appl. geophys., stress decreases linearly to T f over the slip-weakening distance D c. Furthermore, for the inversion, synthetic ground displacement time histories, band-pass filtered between.7 and.5 Hz, are generated using a Green s function propagator method, which is more efficient than the FD method in the case of relatively large faultreceiver distances as in our study. Effect of Initial Parameters on Rupture Propagation Figure 2 illustrates the complexity of the horizontal slip rate generated from heterogeneities in the initial conditions for two sets of the initial stress (T e1 and T e2 ), yield stress (T u1 and T u2 ) and slip-weakening distance (D c1 and D c2 ), compared to a homogeneous reference model. Clearly, small variations in the initial conditions generate large, nonlinear changes in the rupture history. We can generate different rupture histories with very similar initial conditions (for example, T e1 and T e2 ), and similar rupture histories with completely different models (for example, the asperity model with T e1 and the barrier model with D c1 ). Figure 3 depicts significant differences between synthetics generated at the stations YER and for the different initial conditions in Figure 2. Note in particular the strong effects of the initial conditions on the amplitudes in the forward rupture direction (YER) compared to those in the opposite direction (), where the influence of the propagation path between the source and the receiver becomes more significant than the source itself. For this reason, and also due to their smaller amplitudes, stations located in the backward rupture direction are relatively more difficult to use in the rupture inversion. Finally, the observed seismic waveforms are Figure 2 Influence of the initial parameters (top panels) on the horizontal slip rate history for six different heterogeneous models of the Landers earthquake. The other initial parameters for each model are kept constant over the fault. The figure on the left shows a reference model with homogeneous initial parameters.

Vol. 161, 24 Which Dynamic Rupture Parameters can be Estimated 2159 YER Ref 56.4 Ref 6.6 T e1 14.5 T e1 3.3 T e2 13.7 T e2 3.4 T u1 39.8 T u1 5.6 T u2 21.3 T u2 3.6 D c1 19.5 D c1 1.5 D c2 19.9 D c2 3.2 4 s 4 s Figure 3 Horizontal east-west displacement time histories at stations YER and (see Fig. 1) for the models shown in Figure 2. Amplitudes are marked to the right of each trace in centimeters. temporal averages influenced by the entire rupture history. For example, the slip in the epicentral area is controlled mainly by the local stress decrease, whereas the slip toward the end of the fault is influenced by a longer time window of radiation. Similar effects can be seen for the other parameters, the yield stress and the slipweakening distance. In summary, the radiated waves are strongly sensitive to perturbations in the initial dynamic parameters. From Asperity to Barrier Model MADARIAGA and OLSEN (2) identified a non-dimensional parameter that contains all the initial dynamic rupture parameters discussed in this study and controls rupture, j ¼ T 2 e W lðt u T f ÞD c ; ð2þ where W is a characteristic length scale. This parameter was derived from Griffith s criterion for shear faults and measures the ratio of the available strain energy ðu ¼ T 2 e W/2l to the energy release rate ðg ¼ :5ðT u T f ÞD c Þ. For small values of j, rupture does not propagate because the Griffith s criterion is not satisfied. A bifurcation of the rupture occurs when j exceeds a critical value j c, beyond which rupture grows indefinitely. MADARIAGA and OLSEN (2) estimated j c at.7 and.8 for homogeneous asperity and barrier models, respectively. Thus, for models with similar W (for homogeneous models equal to the half width of the fault, see MADARIAGA and OLSEN, 2), j provides a means to relate asperity and barrier models (with T f ¼ MPa) as

216 Sophie Peyrat et al. Pure appl. geophys., Asperity Barrier T e T e T u T u 9 2 MPa Sliprate history 2 s Sliprate history 2 s 4 s 4 s 6 s 6 s 8 s 8 s 1 s 1 s 12 s 12 s 14 s 14 s 3 m/s Figure 4 Comparison between the horizontal slip rate histories for two simple asperity and barrier models. D c is homogeneous and equal to 8 for both models. Rupture is initiated at the circle. j / T e 2 ðx; yþ T u D c Te 2 T u ðx; yþd c Te 2 T u D c ðx; yþ asperity barrier barrier In other words, the model of rupture propagation in asperity and barrier models should be similar when the values of j are identical. However, the problem is more complicated in the heterogeneous case. Figure 4 shows rupture propagation for simple barrier and asperity models derived from Equation 3 assuming constant W and constant D c. It is clear that the rupture propagations for the two models are appreciably different (for example at 6 s), ð3þ

Vol. 161, 24 Which Dynamic Rupture Parameters can be Estimated 2161 Figure 5 Initial conditions and final slip distribution for (a) the asperity model with spatially variable initial stress, (b) the barrier model with spatially variable yield stress, and (c) the barrier model with spatially variable D c. suggesting that the latter assumption is incorrect. Besides, Figure 4 shows smaller differences between the two models further away from the hypocenter. Thus, the values of W could possibly be a function of the distance to the hypocenter (G. BEROZA, personal communication, 2). In any case we expect W to be a representation of the size of the heterogeneities and their correlation length. PEYRAT et al. (21, 22) constructed a barrier model (T u variable) from their asperity model (T e variable) of the 1992 Landers earthquake using Equation (3). Due to the unknown values of W, additional inversion of recorded accelerograms was required to obtain a satisfactory fit between recorded and synthetic accelerograms from the barrier model. In a similar fashion we generate here an additional barrier model with variable D c. Figure 5 shows the initial conditions of these three

2162 Sophie Peyrat et al. Pure appl. geophys., complementary dynamic models. The two barrier models manifest larger and more abrupt variations in their parameter distribution than is the case for the asperity model. We realize that these large values of D c are arguable. Nevertheless, such values of D c and T u are necessary to control rupture propagation in these barrier models. The values of the patches control the path along which rupture propagates for the asperity model and the path where rupture does not propagate for the barrier model. However, the problem of determining the strength of the barriers remains. The straightforward assumption is to assign as large a strength as possible and still promote rupture around these barriers. For example, in a barrier model with variable T u, how large values of the maximum yield stress can be imposed and still retain rupture propagation? However, different values of the maximum T u will generate different rupture histories: the smaller the values of T u the faster the rupture speed and the larger the slip velocity amplitude, if the level of rupture resistance is sufficiently low to allow rupture to propagate. In the case of the barrier model we find very large values of the maximum D c which prevent rupture in these patches but allow rupture to propagate around, complicating the trial and error inversion considerably. Rupture seems more sensitive to small variations in the values of the slip-weakening distance compared to the yield stress. Our three dynamic models generate similar rupture histories (Fig. 6, top) and slip distributions (Fig. 5), i.e., there are no major differences between the asperity model and the barrier models even though the initial conditions are completely different. The only difference is that the barrier model with T u variable generates a higher frequency response due to the higher stress drop. For all models rupture follows a complex path over the fault, showing a confined band of slip propagating unilaterally toward the northwestern part of the fault, before arrest after about 21 s. The complex rupture path and the healing are consequences of the spatial heterogeneities of different initial parameters as originally suggested by BEROZA and MIKUMO (1996). Figure 7 shows a comparison between ground displacement time histories for data and synthetics calculated from the rupture propagation obtained for the three models used in the trial-and-error inversion. The overall waveforms are well reproduced by the synthetics, and the best fits are obtained for stations in the forward rupture direction (YER, BRS, and FTI). Thus we have computed three mechanical models of the rupture propagation which satisfy observed accelerograms. Comparison to Geodetic and Field Data The rupture inversion of the Landers earthquake was carried out using strong motion data only. However, additional observed data may help constrain the rupture model. For example, HERNANDEZ et al. (1999) used simultaneous inversion of interferometric synthetic aperture radar (InSAR), Global Positioning System (GPS), and strong motion data to obtain a kinematic model of the event. Both GPS and

Vol. 161, 24 Which Dynamic Rupture Parameters can be Estimated 2163 Figure 6 Rupture history for the three complementary models of the Landers earthquake. (Top) Each snapshot depicts the horizontal slip rate on the fault at 2 s time intervals. (Bottom) j distributions over the fault. The arrow depicts an average value of j c estimated by MADARIAGA and OLSEN (2) for homogeneous asperity and barrier models. SAR data allow estimates of relative displacements of the ground during an earthquake, without prior knowledge of the location of the event. SAR interferometry was used by MASSONNET et al. (1993) to capture the movements produced by the Landers earthquake. Such data complement strong motion seismic data in their dense spatial sampling. The interferogram was constructed by combining topographic information with SAR images obtained by the ERS-1 satellite before (April 24, 1992) and after (August 7, 1992) the earthquake, covering an area of dimensions 9 km by 11 km (Fig. 8f, top). The SAR interferogram provides a denser spatial sampling (1 m per pixel) than ground-based surveying methods and a better

2164 Sophie Peyrat et al. Pure appl. geophys., North 4 BKR Heterogeneous Te East BKR 4 North Heterogeneous Tu BKR East BKR North 4 BKR Heterogeneous Dc East BKR -4 4 BRN BRN -4 4 BRN BRN -4 4 BRN BRN -4 4 BRS BRS -4 4 BRS BRS -4 4 BRS BRS -4 4 FTI FTI -4 4 FTI FTI -4 4 FTI FTI -4 4 H5 H5-4 4 H5 H5-4 4 H5 H5-4 4 INI INI -4 4 INI INI -4 4 INI INI -4 4 PSA PSA -4 4 PSA PSA -4 4 PSA PSA -4 4-4 4-4 4-4 4 YER YER -4 4 YER YER -4 4 YER YER -4 2 4 6 8 sec 2 4 6 8 sec -4 2 4 6 8 sec 2 4 6 8 sec -4 2 4 6 8 sec 2 4 6 8 data synthetics Figure 7 Comparison between observed ground displacements and those obtained for our inverted dynamic models of the 1992 Landers earthquake. The overall waveforms are well reproduced by the synthetics.

Vol. 161, 24 Which Dynamic Rupture Parameters can be Estimated 2165 precision (3 ) than previous space imaging techniques. The resulting interferogram (Fig. 8f, top) is a contour map of the change of range, i.e., the component of the displacement which points toward the satellite. Each fringe (one interval of color from the red to the blue) corresponds to one cycle, equivalent to 28 mm (half the wavelength of the ERS-1 satellite). In order to test the resolution of InSAR and GPS data to constrain rupture parameters, we computed the three-component area-wide distribution of ground displacement (Figs. 8a c, top) from the slip distribution (Figs. 5a c) for our three dynamic rupture models using OKADA (1985) s formulation. For comparison we show the response of a reference model with constant slip computed as the mean of the distributions generated by the asperity model (Fig. 8d, top), and a model with the slip distribution of the asperity model projected onto three segments (Fig. 8e, top). The overall shape of the fringes for the single-segment asperity and barrier models is well reproduced (Figs. 8a c, top), while the more realistic three-segment model (Fig. 8e, top) shows an improved fit to the data. The agreement between the threesegment synthetic and observed interferograms is satisfactory except for the effects of the M 6.2 Big Bear event which occurred on the same day as the Landers earthquake are not included in the models. Thus, our results show that InSAR data provide a constraint on the geometry of the fault. Furthermore, the shape and amplitude of the fringes for the synthetic interferograms from a constant slip distribution (Fig. 8d, top) are much smoother and smaller, respectively, compared to those generated by the more realistic heterogeneous model. Therefore, the modeling of SAR interferograms provides a constraint on the spatial variation of the slip distribution. We find the same conclusions from similar models of horizontal GPS data (Fig. 8, bottom). Moreover, as shown in Figure 9, the slip distribution along the surface rupture produced by our dynamic models agrees well with the field data of surface displacement (HAUKSSON, 1994; SIEH et al., 1993). In summary, we find it to be important to include seismic, geodetic and field data to obtain the strongest constraints on the dynamic rupture parameters. Discussion and Conclusions We have constructed three well-posed mechanical models, an asperity and two barrier models of the Landers event, which generate synthetics in agreement with strong motion, GPS, InSAR, and field data, confirming an earlier hypothesis that seismic data cannot distinguish between barriers and asperities (see MADARIAGA, 1979). For all models the mode of rupture propagation from the initial asperity patch, the rupture speed, and the healing are critically determined by the distribution of prestress and the friction parameters. Therefore, a critical balance between initial stress and the frictional parameters must be met in order to generate a rupture history which produces synthetics in agreement with data. The barrier and asperity

2166 Sophie Peyrat et al. Pure appl. geophys., Figure 8 (Top) (a e) Synthetic interferograms calculated on a rectangular grid with 2 m spacing for the Landers earthquake. (f) Observed coseismic interferogram covering a 9 by 11 km area measured between April 24 and August 7, 1992 (MASSONNET et al., 1993). Each cycle of the interferogram colors (red to blue) represents 28 mm of ground motion in the direction of the satellite. Black segments depict the fault geometry for each model. (Bottom) (a e) Predicted horizontal GPS displacements calculated for the Landers earthquake depicted by arrows on a regular grid. (f) Observed horizontal displacements at GPS sites and USGS trilateration stations (HUDNUT et al., 1994; MURRAY et al., 1993).

Vol. 161, 24 Which Dynamic Rupture Parameters can be Estimated 2167 6 5 4 Observed surface slip Heterogeneous T e Heterogeneous Tu Heterogeneous D c slip (m) 3 2 1 2 4 6 8 distance along the fault from NW to SE (km) Figure 9 Comparison of simulated surface slip along the fault for the three dynamic models and the observed slip for the Landers earthquake (HAUKSSON, 1994; SIEH et al., 1993). models were related using the non-dimensional parameter j, where we assumed a constant value of the characteristic length scale W (MADARIAGA and OLSEN, 2). In order to examine this assumption we calculated the distributions of j for the three models (Fig. 6, bottom). The distribution of j is somewhat similar for the asperity model and barrier model with variable T u, but differs considerably for the barrier model with variable D c. This result may be due to the nature of the heterogeneous parameters (stress versus distance) in the three models but further supports the view that W varies across the fault for heterogeneous rupture models. W is very likely a measure of the correlation length of the heterogeneities. Future efforts should focus on estimating the (spatially variable?) W that allows j to connect dynamic barrier to asperity rupture models. There are no major differences between the three models of the Landers earthquake in terms of the capacity to reproduce observed data, even though the initial conditions, and therefore their physical interpretation, are completely different. The three models are end-members of a large family of dynamically correct models. It is important to point out that we see no clear indications in the data as to which model is better, and the true rupture for the Landers earthquake is likely a combination of these models. Thus we have demonstrated that the solution of the dynamic problem is not unique, and it may not be possible to separate strength drop and D c using rupture modeling with current bandwidth limitations, in agreement with the conclusions by GUATTERI and SPUDICH (2). However, the inclusion of higher frequency components from high-resolution near-field records of large earthquakes may enable their separation in the future by increasing the resolution of the model. Furthermore, we realize that dynamic parameters are relatively poorly constrained, and additional information on stress drop and the nature of the friction law should be obtained by different methods, such as borehole

2168 Sophie Peyrat et al. Pure appl. geophys., drilling or laboratory experiments. Finally, geodetic and field data may be used to further constrain the rupture propagation parameters in future studies. For example, geodetic data can provide important constraints on the slip distribution and geometry of the fault, in particular in the absence of surface rupture. Acknowledgements This research was supported by grants from the project Mode lisation des séismes et changement d e chelles from the ACI Pre vention des catastrophes naturelles from the Ministe` re de la Recherche, and by NSF (EAR3275), University of California at Sarta Barbara, Los Alamos National Laboratories, and the Southern California Earthquake Center. SCEC is funded by NSF Cooperative Agreement EAR-16924 and USGS Cooperative Agreement 2HQAG8. The SCEC contribution number for this paper is 72, and the ICS contribution number is 556. The computations in this study were carried out on the SUN Enterprise at MRL/ICS and on the facilities of the De partement de Mode lisation Physique et Nume rique of the Institut de Physique du Globe de Paris. REFERENCES BEROZA, G. C. and MIKUMO, T. (1996), Short Slip Duration in Dynamic Rupture in the Presence of Heterogeneous Fault Properties, J. Geophys. Res. 11, 22,449 22,46. DAS, S. and AKI, K. (1977), A Numerical Study of Two-dimensional Spontaneous Rupture Propagation, Geophys. J. Roy. Astr. Soc. 5, 643 668. DUNHAM, E. M., FAVREAU, P., and CARLSON, J. M. (23), A Supershear Transition Mechanism for Cracks, Science 299, 1557 1559. GUATTERI, M. and SPUDICH, P. (2), What Can Strong-motion Data Tell Us about Slip-weakening Faultfriction Laws?, Bull. Seismol. Soc. Am. 9, 98 116. HAUKSSON, E. (1994), State of Stress from Focal Mechanisms Before and After the 1992 Landers Earthquake Sequence, Bull. Seismol. Soc. Am. 84, 917 934. HERNANDEZ, B., COTTON, F., and CAMPILLO, M. (1999), Contribution of Radar Interferometry to a Twostep Inversion of the Kinematic Process of the 1992 Landers Earthquake, J. Geophys. Res. 14, 13,83 13,99. HUDNUT, K. W., BOCK, Y., CLINE, M., FANG, P., FENG, Y., FREYMUELLER, J., GE, X., GROSS, W. K., JACKSON, D., KIM, M., KING, N. E., LANGBEIN, J., LARSEN, S. C., LISOWSKI, M., SHEN, Z. K., SVARC, J., and ZHANG, J. (1994), Co-seismic Displacements of the 1992 Landers Earthquake Sequence, Bull. Seismol. Soc. Am. 84, 625 645. IDA, Y. (1972), Cohesive Force Across the Tip of a Longitudinal-shear Crack and Griffith s Specific Surface Energy, J. Geophys. Res. 77, 3796 385. IDE, S. and TAKEO, M. (1997), Determination of Constitutive Relations of Fault Slip Based on Seismic Wave Analysis, J. Geophys. Res. 12, 27,379 27,391. KANAMORI, H. and STEWART, G. S. (1978), Seismological Aspects of the Guatemala Earthquake of February 4, 1976, J. Geophys. Res. 83, 3427 3434. MADARIAGA, R. (1979), On the Relation Between Seismic Moment and Stress Drop in the Presence of Stress and Strengh Heterogeneity, J. Geophys. Res. 84, 2243 225.

Vol. 161, 24 Which Dynamic Rupture Parameters can be Estimated 2169 MADARIAGA, R., OLSEN, K., and ARCHULETA, R. (1998), Modeling Dynamic Rupture in a 3-D Earthquake Fault Model, Bull. Seismol. Soc. Am. 88, 1182 1197. MADARIAGA, R. and OLSEN, K. B. (2), Criticality of Rupture Dynamics in 3-D, Pure Appl. Geophys. 157, 1981 21. MASSONNET, D., ROSSI, M., CARMONA, C., ADRAGNA, F., PELTZER, G., FEIGL, K., and RABAUTE T. (1993), The Displacement Field of the Landers Earthquake Mapped by Radar Interferometry, Nature 364, 138 142. MURRAY, M. H., SAVAGE, J. C., LISOWSKI, M., and GROSS, W. K. (1993), Coseismic Displacements: 1992 Landers, California, Earthquake, Geophys. Res. Lett. 2, 623 626. OKADA, Y. (1985), Surface Deformation Due to Shear and Tensile Faults in a Half-space, Bull. Seismol. Soc. Am. 75, 1135 1154. OLSEN, K. B., MADARIAGA, R., and ARCHULETA, R. J. (1997), Three-dimensional Dynamic Simulation of the 1992 Landers Earthquake, Science 278, 834 838. PEYRAT, S., MADARIAGA, R., and OLSEN, K. (22), La Dynamique des Tremblements de Terre Vue a` Travers le Se isme de Landers du 28 Juin 1992, C. R. Acad. Sci., Me canique 33, 235 248. PEYRAT, S., OLSEN, K., and MADARIAGA, R. (21), Dynamic Modeling of the 1992 Landers Earthquake, J. Geophys. Res. 16, 26,467 26,482. SIEH, K., JONES, L., HAUKSSON, E., HUDNUT, K., EBERHART-PHILLIPS, D., HEATON, T., HOUGH, S., HUTTON, K., KANAMORI, H., LILJE, A., LINDVALL, S., MCGILL, S. F., MORI, J., RUBIN, C., SPOTILA, J. A., STOCK, J., THIO, H. K., TREIMAN, J., WERNICKE, B., and ZACHARIASEN J. (1993), Nearfield Investigations of the Landers Earthquake Sequence, April to July 1992, Science 26, 171 176. WALD, D. J. and HEATON, T. H. (1994), Spatial and Temporal Distribution of Slip for the 1992 Landers, California, Earthquake, Bull. Seismol. Soc. Am. 84, 668 691. (Received September 27, 22, revised April 25, 23, accepted May 15, 23) To access this journal online: http://www.birkhauser.ch