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

Similar documents
JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 108, NO. B11, 2540, doi: /2002jb002296, 2003

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

Far-reaching transient motions after Mojave earthquakes require broad mantle flow beneath a strong crust

Ground displacement in a fault zone in the presence of asperities

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

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

3D Finite Element Modeling of fault-slip triggering caused by porepressure

Synthetic Seismicity Models of Multiple Interacting Faults

Earthquake and Volcano Deformation

Effect of an outer-rise earthquake on seismic cycle of large interplate earthquakes estimated from an instability model based on friction mechanics


DETAILS ABOUT THE TECHNIQUE. We use a global mantle convection model (Bunge et al., 1997) in conjunction with a

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

High-Harmonic Geoid Signatures due to Glacial Isostatic Adjustment, Subduction and Seismic Deformation

SUPPLEMENTARY INFORMATION

Influence of anelastic surface layers on postseismic

Kinematics of the Southern California Fault System Constrained by GPS Measurements

Rheology III. Ideal materials Laboratory tests Power-law creep The strength of the lithosphere The role of micromechanical defects in power-law creep

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

D DAVID PUBLISHING. Deformation of Mild Steel Plate with Linear Cracks due to Horizontal Compression. 1. Introduction

Secondary Project Proposal

Asish Karmakar 1, Sanjay Sen 2 1 (Corresponding author, Assistant Teacher, Udairampur Pallisree Sikshayatan (H.S.), Udairampur, P.O.

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

SLIP DISTRIBUTION FOR THE 2004 SUMATRA-ANDAMAN EARTHQUAKE CONSTRAINED BY BOTH GPS DATA AND TSUNAMI RUN-UP MEASUREMENTS

Data Repository Hampel et al., page 1/5

Originally published as:

Deformation cycles of great subduction earthquakes in a viscoelastic Earth

M 7.0 earthquake recurrence on the San Andreas fault from a stress renewal model

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

Shear Stresses and Displacement for Strike-slip Dislocation in an Orthotropic Elastic Half-space with Rigid Surface

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

Time-Dependent Viscoelastic Stress Transfer and Earthquake Triggering

Differentiating earthquake tsunamis from other sources; how do we tell the difference?

Elizabeth H. Hearn modified from W. Behr

Three-dimensional viscoelastic finite element model for postseismic deformation of the great 1960 Chile earthquake

Aftershocks and Pore Fluid Diffusion Following the 1992 Landers Earthquake

Expansion of aftershock areas caused by propagating post-seismic sliding

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

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

The Mechanics of Earthquakes and Faulting

overlie the seismogenic zone offshore Costa Rica, making the margin particularly well suited for combined land and ocean geophysical studies (Figure

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

Postseismic Deformation from ERS Interferometry

Lecture 20: Slow Slip Events and Stress Transfer. GEOS 655 Tectonic Geodesy Jeff Freymueller

FRICTIONAL HEATING DURING AN EARTHQUAKE. Kyle Withers Qian Yao

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

Co-seismic Gravity Changes Computed for a Spherical Earth Model Applicable to GRACE Data

Long-term Crustal Deformation in and around Japan, Simulated by a 3-D Plate Subduction Model

Measurements in the Creeping Section of the Central San Andreas Fault

Modeling of co- and post-seismic surface deformation and gravity changes of M W 6.9 Yushu, Qinghai, earthquake

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

The Effect of Elastic Layering on Inversions of GPS Data for Coseismic Slip and Resulting Stress Changes: Strike-Slip Earthquakes

Coupled afterslip and viscoelastic flow following the 2002

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

The Earthquake Cycle Chapter :: n/a

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

CHARACTERIZING EARTHQUAKE SLIP MODELS FOR THE PREDICTION OF STRONG GROUND MOTION

A viscoelastic model of interseismic strain concentration in Niigata-Kobe Tectonic Zone of central Japan

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

Scaling relations of seismic moment, rupture area, average slip, and asperity size for M~9 subduction-zone earthquakes

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

Coseismic slip distribution of the 1946 Nankai earthquake and aseismic slips caused by the earthquake

arxiv: v1 [physics.geo-ph] 30 Apr 2013

Verification of the asperity model using seismogenic fault materials Abstract

A THEORETICAL MODEL FOR SITE COEFFICIENTS IN BUILDING CODE PROVISIONS

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

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

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

State of Stress in Seismic Gaps Along the SanJacinto Fault

Deformation of a layered half-space due to a very long tensile fault

Splay fault and megathrust earthquake slip in the Nankai Trough

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

EARTHQUAKE LOCATIONS INDICATE PLATE BOUNDARIES EARTHQUAKE MECHANISMS SHOW MOTION

Graves and Pitarka Method

Source rupture process of the 2003 Tokachi-oki earthquake determined by joint inversion of teleseismic body wave and strong ground motion data

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

Frictional Properties on the San Andreas Fault near Parkfield, California, Inferred from Models of Afterslip following the 2004 Earthquake

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

Evidence for postseismic deformation of the lower crust following the 2004 Mw6.0 Parkfield earthquake

Earthquake Doublet Sequences: Evidence of Static Triggering in the Strong Convergent Zones of Taiwan

Three-dimensional numerical simulation on the coseismic deformation of the 2008 M S 8.0 Wenchuan earthquake in China

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

Tectonic Seismogenic Index of Geothermal Reservoirs

} based on composition

Continuum mechanism: Stress and strain

Spectral Element simulation of rupture dynamics

Geophysical Journal International

Seismic and aseismic processes in elastodynamic simulations of spontaneous fault slip

Characterizing Earthquake Rupture Models for the Prediction of Strong Ground Motion

Two Neighbouring Strike Slip Faults and Their Interaction

Study megathrust creep to understand megathrust earthquakes

3D MODELING OF EARTHQUAKE CYCLES OF THE XIANSHUIHE FAULT, SOUTHWESTERN CHINA

Earthquake and Volcano Clustering at Mono Basin (California)

SDSU Module Kim Olsen and Rumi Takedatsu San Diego State University

Three Dimensional Simulations of Tsunami Generation and Propagation

The Size and Duration of the Sumatra-Andaman Earthquake from Far-Field Static Offsets

Inferring fault strength from earthquake rupture properties and the tectonic implications of high pore pressure faulting

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

SONGS SSC. Tom Freeman GeoPentech PRELIMINARY RESULTS

Finite element modelling of fault stress triggering due to hydraulic fracturing

Transcription:

Homogeneous vs. realistic heterogeneous material-properties in subduction zone models: Coseismic and postseismic deformation T. Masterlark 1, C. DeMets 2, H.F. Wang 2, O. S nchez 3, and J. Stock 4 1 US Geological Survey, EROS Data Center, Raytheon, Sioux Falls, SD 57198; masterlark@usgs.gov 2 University of Wisconsin-Madison, Madison, WI 53706; chuck@geology.wisc.edu, wang@geology.wisc.edu 3 Instituto de Geof sica, UNAM, M xico D. F., M xico; osvaldo@ollin.igeofcu.unam.mx 4 California Institute of Technology, Pasadena, CA 91125; jstock@gps.caltech.edu 1 Introduction Three-dimensional finite-element models (FEM)s are capable of simulating tectonic deformation during all phases of the seismic cycle because they allow for heterogeneous material property distributions, complicated boundary condition and loading specifications, and contact surface interactions. These models require a dislocation source to drive the coseismic response and the subsequent poroelastic and viscoelastic relaxation. Analytical solutions to compute Green s functions for displacement due to a dislocation are readily available [e.g., Okada, 1992]. These solutions, which include homogeneous elastic halfspace (HEHS) assumptions, are often used in inverse methods to solve for dislocation distributions based on observed coseismic deformation. Alternative methods for generating displacement Green s functions, allowing for heterogeneous material property distributions, require half-space boundary conditions [Du et al., 1997; Savage, 1998]. Using the 1995 Colima-Jalisco earthquake as an example, we compute FEM-generated Green s functions for both drained and undrained HEHS models and for a system with spatially varying material properties characteristic of a subduction zone. Dislocation Figure 1. Coseismic horizontal GPS displacements for the 1995 (M w =8.0) Colima-Jalisco earthquake. The rupture surface intersects the Middle America trench and dips to the northeast. The dashed rectangle is the surface projection of the rupture. Predictions from G C loaded with s A significantly overestimate the deformation magnitudes of the coastal sites. G C s A and G C s C are described in the text. distributions are estimated from inversion of GPS displacements. We then estimate coseismic and postseismic deformation prediction errors introduced by homogeneous material property assumptions and the sensitivity to drained versus undrained conditions. The 9 October 1995 (M w=8.0) Colima-Jalisco earthquake, which ruptured the Rivera-North American plate subduction interface (Figure 1), was the first significant rupture of the Middle America trench northwest of the Manzanillo trough since the 3 June 1932 (M w =8.2) and 18 June 1932 (M w =7.8) earthquakes [Singh et al., 1985]. Inversions, with HEHS model assumptions, of threedimensional coseismic displacements from 11 nearby GPS sites suggest that the seismic moment release was concentrated in two regions, one near the northwest edge of the Manzanillo trough and the other 80-120 km farther northwest [Melbourne et al., 1997, Hutton et al., 2001], in accord with seismologic results [Mendoza and Hartzel, 1999. 2 Assumptions and Techniques: FEM FEMs in this study were constructed with ABAQUS [HKS, Inc., 2000], a commercial finite-element code that allows for poroelastic and viscoelastic material properties and contact surface interactions. A three-dimensional FEM was designed to simulate the subduction zone along the Middle America trench (Figure 2). The 28-km-thick continental crust [Pardo and Suarez, 1995] of the North American plate consists of a 16-km-thick poroelastic upper crust overlying a 12-km-thick linear viscoelastic lower crust. The oceanic crust of the Rivera plate is assumed to be 6 km thick. Poroelastic effects are neglected in the oceanic crust because they are poorly constrained by the lack of

Figure 2. FEM configuration, FEM C. The problem domain is tessellated into 24,750 three-dimensional elements. The expanded ~70 km-thick portion of the near-field region displays the heterogeneous material property distribution and subduction zone geometry. The continental lithosphere is separated into poroelastic upper crust, viscoelastic lower crust, and elastic upper mantle layers. The oceanic lithosphere includes the elastic crust overlying the elastic upper mantle. Boundary and initial conditions are discussed in the text. offshore GPS displacements. The upper mantle extends from the base of the crust in both plates to a depth of about 200 km and is treated as an elastic material. Zero displacement is specified along the lateral boundaries and base of the problem domain. The top of the problem domain is an elastic free surface. The boundaries of the poroelastic upper crust are no-flow surfaces. Parameters chosen are as follows: poroelastic and drained elastic properties for Westerly Granite [Wang, 2000] are used for the poroelastic upper crust and viscoelastic lower crust layers respectively. The bulk hydraulic diffusivity of the upper crust is 10 2 m 2 s -1 [Nur and Walder, 1992; Masterlark, 2000] and the lower crust viscosity is 5 10 18 Pa s [Deng et al., 1998; Masterlark, 2000]. We use typical elastic properties for oceanic crust and mantle rock [Turcotte and Schubert, 1982]. The fault is a convex, deformable contact surface divided into subfaults that measure 20 10 km along-strike and down-dip respectively [Hutton et al, 2001]. Contact-node pairs are located at the center of each patch and along the top edge of the fault patches that intersect the free surface. The two lithospheric plates are welded together along the down-dip and along-strike extensions of the seismogenic portion of the interface. Initial stress and fluid-pressure conditions are geostatic. The three mechanical systems considered are FEM A : drained HEHS, FEM B : undrained HEHS, and FEM C : a heterogeneous material property distribution and poroelastic upper crust. For the two HEHS models, upper crust material properties are specified throughout the system. Drained conditions imply fluid-pressure does not change, a condition assumed in the vast majority of studies of active faults including previous solutions for dislocation distributions of the 1995 Colima-Jalsco earthquake [Melbourne et al., 1997; Mendoza and Hartzel, 1999; Hutton et al., 2001]. Undrained conditions exist in the poroelastic upper crust immediately after a sudden dislocation because stress is transferred throughout the system much faster than fluids can flow [Wang, 2000]. The forward solution for displacements due to a dislocation distribution in an elastic material is a linear system of equations G s = d for an a priori fault geometry, where G is the matrix of displacement Green s functions, s is a vector of dislocations for contact-node pairs, and d is a vector of displacements. Symbolically, the coefficient G ij is a displacement component at location j due to a unit dislocation of contact-node pair i. For the case of the 1995 Colima- Table 1. Weighted least-squares misfit Model Configuration χ 2 Hutton et al. [2000] drained HEHS 166 G A s A drained HEHS 171 G B s B undrained HEHS 171 G C s C undrained heterogeneous 164 G C s A inconsistent load 9000 G C s B inconsistent load 8000 Jalisco earthquake, we consider reverse-slip only [Melbourne et al., 1997, Hutton et al., 2001]. We inverted the 11 measured threedimensional coseismic displacements from Hutton et al. [2001] to obtain the mixeddetermined solution for s using damped leastsquares methods. A weighting matrix is applied to the data vector to account for uncertainties [Menke, 1989]. Models FEM A, FEM B, and FEM C generate G A, G B, and G C, which are used to obtain dislocation distributions s A, s B, and s C respectively (Figure 3). The weighted least-squares misfits, χ 2, from the three models are similar to those determined by Hutton et al. [2001] (Table 1). Dislocation magnitudes determined for an undrained material will be lower than for a drained material because undrained material properties are stiffer than their drained counterparts. The heterogeneous model (FEM C ), which simulates slip along the deformable interface between oceanic and continental crust, is the stiffest of the three cases we

considered due to the undrained upper crust and oceanic crust and mantle properties. Predicted seismic moments are 6.9, 6.7, and 6.5 10 20 N m for s A, s B, and s C respectively, near the mid-point of seismologic estimates. 3 Results 3.1 Coseismic Predictions There are substantial differences between coseismic predictions from FEM C, loaded with s C (the appropriate load) and either s A (Figure 1) or s B. Predictions from both s A and s B but using the coefficient matrix G C overestimate displacements for the coastal GPS sites by as much as 250 mm, much more than the ~6 mm displacement uncertainty. Weighted least-squares misfits from these two models are more than an order of magnitude larger than for displacements predicted by s C and using G C (Table 1). These differences are expected based on the inconsistency between the model used to obtain the dislocation distribution and the one used to predict the displacements. These differences are also found in previous studies that used numerical methods to estimate the importance of heterogeneity on coseismic deformation predictions [e.g., Eberhart-Phillips and Stuart, 1992; Wald and Graves, 2001]. Figure 3. Coseismic dislocation distributions. The magnitudes of predicted reverse-slip dislocation along the rupture are shown for dislocation distributions s A, s B, and s C respectively. The contour interval is 2 meters. Although the distributions are similar, they are not interchangeable among the models from which they were derived. The predicted seismic moment for s A (derived from the drained HEHS model) overestimates the seismic moment of s C (derived from the undrained model with heterogeneous material properties) by about 6 percent. 3.2 Postseismic Predictions The coseismic response of the model to the dislocation load represents the initial conditions for the postseismic model. Fluid-pressure in the upper crust and shear stress in the lower crust, initiated by fault slip, drive poroelastic and viscoelastic relaxation. For a time step of 5 years after the earthquake, which allows for both poroelastic and viscoelastic relaxation contributions, FEM C loaded with s C predicts a horizontal displacement pattern of convergence toward the rupture and subsidence nearly everywhere onshore [Masterlark et al., 2001]. Differences between postseismic horizontal displacements predicted by FEM C, loaded with s C versus s A, can be more than an order of magnitude greater than typical uncertainties (~6 mm) in GPS displacements. Although differences in vertical predictions are similar in magnitude, the uncertainties in vertical GPS displacements are much higher (~15 mm). The root-mean-squared-error in postseismic predictions, with respect to FEM C loaded with s C, for locations corresponding to the 11 GPS sites is 6 mm and 5 mm for FEM C loaded with s A and s B respectively. The comparisons thus far represent the sensitivity to heterogeneous material properties. The two homogeneous FEMs and their dislocation distributions are used to estimate sensitivity to poroelastic effects. The prediction difference between the initially undrained HEHS (FEM B ), loaded with s B versus s A, approximates the error introduced by using a dislocation distribution, determined for a drained condition assumption, in

a model that includes an undrained condition assumption. In this case, the maximum differences in the postseismic horizontal and vertical displacements are 18 mm and 13 mm respectively. The horizontal differences exceed the uncertainties in the GPS displacements. The dislocation distributions derived from HEHS models that assume drained conditions should thus not be used to drive models of postseismic poroelastic deformation [Bosl and Nur, 1998; Peltzer et al., 1998; Masterlark, 2000]. 4 Conclusions For the 1995 Colima-Jalisco M w =8.0 earthquake, the errors introduced in both coseismic and postseismic deformation predictions by including HEHS assumptions exceed by an order of magnitude the estimated uncertainties in the GPS displacements. HEHS assumptions are unnecessary, except for simple first-order approximations, because FEM-generated Green s functions allow for realistic heterogeneous material property distributions, complicated boundary condition and loading specifications, and contact surface interactions. Furthermore, the method we propose allows for additional complexities within the capability of FEM methods. For example, material property anisotropy and topographic effects could be included in the FEM-generated Green s functions. Results of this study are particularly relevant to seismic hazard assessments. Quantitative earthquake coupling analyses require realistic models and assumptions [Masterlark and Wang, 2000]. Transient deformational models can be used to quantify the evolution of quasi-static Mohr-Coulomb frictional states along faults [Masterlark, 2000]. Because these transient models are calibrated to observed displacements, errors introduced from HEHS assumptions should be avoided. References Bosl, W., and A. Nur, Numerical Simulation of Postseismic Deformation due to Pore Fluid Diffusion, in Poromechanics, Edited by J.-F. Thimus, Y. Abousleiman, A.H.-D. Cheng, O. Coussy, and E. Detournay, pp. 23-28, Balkema, Rotterdam, 1998. Deng, J., M. Gurnis, H. Kanamori, and E. Hauksson, Viscoelastic Flow in the Lower Crust after the 1992 Landers, California, Earthquake, Science, 282, 1601-1772, 1998. Du., Y., P. Segall, and H. Gao, Quasi-Static Dislocations in Three-Dimensional Inhomogeneous Media, Geophys. Res. Lett., 24, 2347-2350, 1997. Eberhart-Phillips, D., and W.D. Stuart, Material Heterogeneity Simplifies the Picture: Loma Prieta, Bull. Seism. Soc., 82, 1964-1968, 1992. Hutton, W., C. DeMets, O. Sanchez, G. Suarez, and J. Stock, Slip Kinematics and Dynamics During and After the 9 October 1995 Mw=8.0 Colima-Jalisco Earthquake, Mexico, from GPS Geodetic Constraints, Geophys. Jour. Int., accepted for publication, 2001. Masterlark, T.L., Regional Fault Mechanics Following the 1992 Landers Earthquake, Ph.D. Thesis, 83 pp., University of Wisconsin-Madison, Madison, Wisconsin, 2000. Masterlark, T., and H.F. Wang, Poroelastic Coupling Between the 1992 Landers and Big Bear Earthquakes, Geophys. Res. Lett., 27, 3647-3650, 2000. Masterlark, T., C. DeMets, H.F. Wang, O. S nchez, and J. Stock, Homogeneous vs. heterogeneous subduction zone models: Coseismic and postseismic deformation, Geophys. Res. Lett., accepted for publication, August 2001. Melbourne, T., I. Carmichael, C. DeMets, K. Hudnut, O. Sanchez, J. Stock, G. Suarez, and F. Webb, The Geodetic Signature of the M8.0 Oct. 9, 1995, Jalisco Subduction Earthquake, Geophys. Res. Lett,, 24, 715-718, 1997. Mendoza, C., and S. Hartzell, Fault-Slip Distribution of the 1995 Colima-Jalisco, Mexico, Earthquake, Bull. Seism. Soc., 89, 1338-1344, 1999. Menke, W., Geophysical Data Analysis: Discrete Inverse Theory, 289 pp., Academic Press Inc., San Diego, Calif., 1989. Nur, A. and J. Walder, Hydraulic Pulses in the Earth s Crust, in Fault Mechanics and Transport Properties of Rocks, Edited by B. Evans and T.F. Wong, pp. 461-473, Academic Press, London, 1992. Okada, Y., Internal Deformation due to Shear and Tensile Faults in a Half-Space, Bull. Seism. Soc., 82, 1018-1040, 1992. Pardo, M., and Su\ arez, Shape of the Subducted Rivera Plate in Southern Mexico: Seismic and Tectonic Implications, J. Geophys. Res., 100, 12357-12373, 1995. Peltzer, G., P. Rosen, F. Rogez, and K. Hudnut, Postseismic Rebound Along the Landers 1992 Earthquake Surface Rupture, J. Geophys. Res., 103, 30131-30145, 1998. Savage, J.C., Displacement field for an edge dislocation in a layered half-space, J. Geophys. Res., 103, 2439-2446,

1998. Singh, S. K., L. Ponce, and S.P. Nishenko, The Great Jalisco, Mexico, Earthquakes of 1932: Subduction of the Rivera Plate, Bull. Seismol. Soc. Am., 75, 1301-1313, 1985. Turcotte, D.L., and G.J. Schubert, Geodynamics, 450 pp., John Wiley \& Sons, New York, 1982. Wald, D.J., and R.W. Graves, Resolution Analysis of Finite Fault Source Inversion using One- and Three-Dimensional Green s Functions 2. Combining Seismic and Geodetic Data, J. Geophys. Res., 106, 8767-8788, 2001. Wang, H.F., Theory of Linear Poroelasticity: with Applications to Geomechanics and Hydrogeology, 287 pp., Princeton University Press, Princeton, New Jersey, 2000.