Postseismic deformation of the Andaman Islands following the 26 December, 2004 Great Sumatra-Andaman Earthquake

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

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

Elizabeth H. Hearn modified from W. Behr

Ground displacement in a fault zone in the presence of asperities

Deformation cycles of great subduction earthquakes in a viscoelastic Earth

Azimuth with RH rule. Quadrant. S 180 Quadrant Azimuth. Azimuth with RH rule N 45 W. Quadrant Azimuth

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

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

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

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

Constraints on 2004 Sumatra Andaman earthquake rupture from GPS measurements in Andaman Nicobar Islands

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

Earthquake and Volcano Deformation

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

GPS measurements of postseismic deformation in the Andaman- Nicobar region following the giant 2004 Sumatra-Andaman earthquake

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

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

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

Kinematics of the Southern California Fault System Constrained by GPS Measurements

RELOCATION OF LARGE EARTHQUAKES ALONG THE SUMATRAN FAULT AND THEIR FAULT PLANES

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

Coseismic Slip Distributions of the 26 December 2004 Sumatra Andaman and 28 March 2005 Nias Earthquakes from GPS Static Offsets

What is the LAB Dynamically: Lithosphere and Asthenosphere Rheology from Post-loading Deformation

Supplementary Material

} based on composition

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

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

Seismic and aseismic processes in elastodynamic simulations of spontaneous fault slip

Coupled afterslip and viscoelastic flow following the 2002

The Earthquake of Padang, Sumatra of 30 September 2009 scientific information and update

Variability of earthquake nucleation in continuum models of rate-and-state faults and implications for aftershock rates

Evidence for and implications of a Bering plate based on geodetic measurements from the Aleutians and western Alaska

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

Separating Tectonic, Magmatic, Hydrological, and Landslide Signals in GPS Measurements near Lake Tahoe, Nevada-California

News Release December 30, 2004 The Science behind the Aceh Earthquake

Originally published as:

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

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

The Earthquake Cycle Chapter :: n/a

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

Depth (Km) + u ( ξ,t) u = v pl. η= Pa s. Distance from Nankai Trough (Km) u(ξ,τ) dξdτ. w(x,t) = G L (x,t τ;ξ,0) t + u(ξ,t) u(ξ,t) = v pl

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

Seismotectonics of intraplate oceanic regions. Thermal model Strength envelopes Plate forces Seismicity distributions

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

Secondary Project Proposal

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

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

1.3 Short Review: Preliminary results and observations of the December 2004 Great Sumatra Earthquake Kenji Hirata

Can geodetic strain rate be useful in seismic hazard studies?

The problem (1/2) GPS velocity fields in plate boundary zones are very smooth. What does this smoothness hide?

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

Afterslip, slow earthquakes and aftershocks: Modeling using the rate & state friction law

Using deformation rates in Northern Cascadia to constrain time-dependent stress- and slip-rate on the megathrust

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

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

SUPPLEMENTARY INFORMATION

Report on Banda Aceh mega-thrust earthquake, December 26, 2004

Afterslip and viscoelastic relaxation following the 1999 M 7.4 İzmit earthquake from GPS measurements

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

Synthetic Seismicity Models of Multiple Interacting Faults

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

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

SUPPLEMENTARY INFORMATION

Izmit earthquake postseismic deformation and dynamics of the North Anatolian Fault Zone

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 112, B10402, doi: /2007jb004928, 2007

Electronic supplement for Forearc motion and deformation between El Salvador and Nicaragua: GPS, seismic, structural, and paleomagnetic observations

Elastic Rebound Theory

Sendai Earthquake NE Japan March 11, Some explanatory slides Bob Stern, Dave Scholl, others updated March

Time Dependence of Postseismic Creep Following Two Strike-Slip Earthquakes. Gerasimos Michalitsianos

Measurements in the Creeping Section of the Central San Andreas Fault

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

COULOMB STRESS CHANGES DUE TO RECENT ACEH EARTHQUAKES

INGV. Giuseppe Pezzo. Istituto Nazionale di Geofisica e Vulcanologia, CNT, Roma. Sessione 1.1: Terremoti e le loro faglie

Seismic Activity near the Sunda and Andaman Trenches in the Sumatra Subduction Zone

Report on Banda Aceh mega-thrust earthquake, December 26, 2004

Earthquake distribution is not random: very narrow deforming zones (= plate boundaries) versus large areas with no earthquakes (= rigid plate

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

Principles of the Global Positioning System Lecture 24

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

Geologic Structures. Changes in the shape and/or orientation of rocks in response to applied stress

Does Aftershock Duration Scale With Mainshock Size?

A MODEL OF PLATE CONVERGENCE IN SOUTHWEST JAPAN, INFERRED FROM LEVELING DATA ASSOCIATED WITH THE 1946 NANKAIDO EARTHQUAKE

Post-seismic motion following the 1997 Manyi (Tibet) earthquake: InSAR observations and modelling

ESTIMATES OF HORIZONTAL DISPLACEMENTS ASSOCIATED WITH THE 1999 TAIWAN EARTHQUAKE

Influence of anelastic surface layers on postseismic

Segmentation in episodic tremor and slip all along Cascadia

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

Title. Author(s)Heki, Kosuke. CitationScience, 332(6036): Issue Date Doc URL. Type. File Information. A Tale of Two Earthquakes

Using short-term postseismic displacements to infer the ambient deformation conditions of the upper mantle

MAGMATIC, ERUPTIVE AND TECTONIC PROCESSES IN THE ALEUTIAN ARC, ALASKA

The Mechanics of Earthquakes and Faulting

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

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

Gravity Tectonics Volcanism Atmosphere Water Winds Chemistry. Planetary Surfaces

Structural context of the great Sumatra-Andaman Islands earthquake

The importance of the South-American plate motion and the Nazca Ridge subduction on flat subduction below South Peru

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

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

Geophysical Journal International

Qualitative modeling of earthquakes and aseismic slip in the Tohoku-Oki area. Nadia Lapusta, Caltech Hiroyuki Noda, JAMSTEC

Transcription:

GEOPHYSICAL RESEARCH LETTERS, VOL.???, XXXX, DOI:1.129/, Postseismic deformation of the Andaman Islands following the 26 December, 24 Great Sumatra-Andaman Earthquake J. Paul, 1 A. R. Lowry, 2 R. Bilham, 3 S. Sen, 4 and R. Smalley, Jr. 1 Two years after the Great Sumatra-Andaman earthquake the 3.1 m WSW coseismic displacement at Port Blair, Andaman Islands, had increased by 32 cm. Postseismic uplift initially exceeded 1 cm per week and decreased to <1 mm/week. By 27 points near Port Blair had risen more than 2 cm, a 24% reversal of coseismic subsidence. Uplift at eight GPS sites suggests a gradual eastward shift of the coseismic neutral axis separating subsidence from uplift. Simulations of the GPS postseismic displacements as viscoelastic relaxation of coseismic stress change and as slip on the plate interface indicate that slip down-dip of the seismic rupture dominates near-field deformation during the first two years. Postseismic slip beneath the Andaman Islands released moment equivalent to a magnitude M w 7.5 earthquake, and the distribution suggests deep slip in the stable frictional regime accelerated to catch up to the coseismic rupture. 1. Introduction The December 26 th 24, M w=9.3 Sumatra-Andaman earthquake ruptured 16 km of subduction thrust at the east boundary of the Indian plate. In addition to several meters of southwestward coseismic displacement (Figure 1- inset), Global Positioning System (GPS) instruments measured 6 cm of coseismic uplift at Diglipur in the Northern Andaman Islands, 84 cm subsidence at Port Blair, 7 cm subsidence at Havelock island, and 34 cm uplift at Hut Bay [Freymueller et al., 27]. Remote sensing and visual inspection of shoreline changes provide evidence for both uplift and subsidence, ranging from +4 to 7 cm, which Meltzner et al. [26] characterize as separated by a hinge line. Others have used the term pivot line. Because the coseismic surface flexure is not truly a planar tilt, in this article we refer to this line of zero uplift as a neutral axis. Postseismic processes discussed here have translated this line eastward (Figure 1). We collected GPS measurements of postseismic deformation at eleven sites in the Andaman Islands (Figures 1 and 2). We initiated continuous GPS recording at Port Blair () three weeks after the earthquake, Havelock () in January 27, and on Little Andaman (HUTB) in December 26. Other sites were occupied for periods of several 1 Centre for Earthquake Research and Information, 3876 Central Ave, Memphis, TN 38152 USA. 2 Department of Geology, Utah State University, Logan, UT 84322-455, USA. 3 Department of Geological Sciences, University of Colorado, Boulder, CO 839-399, USA. 4 Society for Andaman and Nicobar Ecology (SANE), Middle Point, PO Box 63, Port Blair 74411, India. Copyright 27 by the American Geophysical Union. 94-8276/7/$5. days at intervals of several months. The data afford sufficient spatial and temporal sampling to characterize nearfield postseismic deformation. Three deformation processes are likely candidates for postseismic response to coseismic stress change: (1) poroelastic relaxation moderated by flow of interstitial fluids [Peltzer et al., 1998]; (2) viscoelastic relaxation of the mantle [Rundle, 1978]; and (3) aseismic slip in velocitystrengthening frictional conditions [Tse and Rice, 1986]. Surface deformation caused by these processes can look similar, such that the effects of one process may be mismodeled using the physics of another [Thatcher and Rundle, 1979; Fialko, 25]. Laboratory deformation experiments leave little doubt that all three processes contribute to transient deformation after earthquakes, but the relative contribution of each should depend on temporal and spatial scales of measurement. Where data are adequately sampled in both time and space, all three processes may be required to fit the observations [Freed et al., 26]. However examples of great earthquake postseismic deformation sampled densely in both time and space are few. Consequently questions remain as to the roles of these three processes in the earthquake cycle, and even whether they play a similar role on all fault zones or in subsequent events on the same fault zone. In this article we combine campaign and continuous GPS data to examine end-member models of postseismic deformation. The first two years of near-field Andaman deformation predominantly reflects postseismic slip downdip of the coseismic rupture, consistent with simulations of frictional slip dynamics. 2. Data and Analysis GPS measurements were initiated at six sites (,,,, and ) within two to six weeks after the earthquake. We collected three to five epochs of data at seven sites during the first two years (Figure 2). All of the campaign GPS monuments consist of steel pins drilled into rock. At continuous sites, monuments consist of 4 m concrete pillars that rise to 2 m above ground. At 2 m depth, the pillars are anchored to a 3 cm thick, 4 4 m reinforced concrete foundation. Data analysis at the University of Memphis used GAMIT/GLOBK. Twelve regional IGS sites were used to determine the ITRF2 reference frame. In the first two years beginning 2 days after the mainshock, station moved 7.5 cm south, 31 cm west and rose 23 cm (Figure 2). All of the campaign sites exhibit uplift and SW to WSW motion, but magnitudes vary by up to 2% and azimuths differ by as much as 36. Motion at Mount Hariot () and Rangachang () is very similar to that at nearby. Wandoor () and Rutland () have smaller, more westerly displacements 1 than sites further east. and continuous site both exhibit a large southward component of motion. Ramnagar () moved the least among the well-sampled sites. Prior to the earthquake, four epochs of campaign GPS measurement were collected at a site near in 1996 1999 (CARO, subsequently destroyed). Our reanalysis of

X - 2 PAUL ET AL.: ANDAMAN POSTSEISMIC DEFORMATION CARO data indicates 9.3±1.8 mm/yr right-lateral oblique convergence with the Indian plate [Paul et al., 21]. How this relates to India Burma relative plate motion depends on unknown coupling of the two, and the true plate convergence rate is probably more than twice this [e.g., Socquet, 26]. The Andamans are approximately 1 km from the trench, and GPS measurements of coseismic rupture suggest that Port Blair lies slightly east of the downdip terminus (Figure 1-inset). 3. Modeling 3.1. Exponential Fit of GPS Time Series To compare the mix of continuous and campaign data with model predictions, we first estimate transient displacements at each site. is the best-sampled GPS site (Figure 2), and the coordinate time series x(t) there suggests an exponential decay of the form ( )] x(t) = x + V t A T t [1 exp (1) τ in which V is constant (interseismic) velocity, T is time of the Great Sumatra/Andaman earthquake, τ is characteris- ' 11 3' 92 3' 14 13.5 11 1.5 -.5 -.4 -.3 -.2 -.1..1.2.3.4.5 Coseismic Shoreline Uplift (m) Coseismic 1 m Postseismic 1 m GPS & Tide Gauge INTV Coseismic Vertical SMTH 1 93 ' Postseismic Slip (m) 5 cm HUTB. 2.5. 2.5 5. 7.5 1. 12.5 15. 17.5 2. Coseismic Slip (m) Figure 1. Andaman Islands uplift. Black inverted triangles are GPS sites measured for postseismic displacement in this study. Circles are remote-sensed coseismic shoreline uplift from Meltzner et al. [26]; solid line is their estimate of neutral axis separating western uplift from eastern subsidence. Gray vectors are coseismic vertical from GPS and tide gauge. Dashed line approximates current neutral axis summing coseismic and postseismic. Inset: Best-fit models of Andaman coseismic fault slip from GPS data summarized in Freymueller et al. [27] (red with black vectors), and postseismic slip from GPS data described here (blue with gray vectors). tic timescale of the decay, and A is transient displacement in the limit as t. Interseismic velocity is assumed to be V =(3.59,3.2,.) cm/yr at all of the sites, i.e., velocity at CARO in the ITRF-2 frame of the postseismic data. This assumption will be correct (within uncertainties) at nearby, but interseismic slip modeling in other subduction zones [e.g., Lowry, 26] suggests this introduces errors of up to 5 mm/yr at the other sites (i.e., <3% of the 35 5 cm transient displacements). Displacement terms A are estimated by weighted least-squares fit of the data to eqn 1. The choice of τ minimizing the weighted L 2-norm misfit is determined by grid search. Simultaneous parameter inversion of all the Andaman GPS time series yields an exponential decay timescale τ=.82+.7/-.6 years at 95% confidence. Estimates of horizontal and vertical transient displacement are depicted in Figure 3. Here we show displacements estimated for the period T 27. (i.e., we multiply A by.914 to represent partially completed decay). East (m) North (m) Up (m).4.2. -.2 -.4.4.2. -.2 -.4.4.2. -.2 -.4 2 5 1 2 5 Days After the Earthquake Figure 2. Daily coordinate solutions at Andaman GPS sites versus log-scaled time after the earthquake. Dashed lines are best-fit models of exponential decay plus interseismic velocity.

PAUL ET AL.: ANDAMAN POSTSEISMIC DEFORMATION X - 3 3.2. Viscoelastic Relaxation We used VISCO1D [Pollitz, 1997] to model viscoelastic response of a self-gravitating, layered spherical earth. Sumatra/Andaman coseismic slip was approximated as three fault planes dipping 15 to 4 km depth. Coseismic slip was approximated from seismic and geodetic estimates [e.g., Lay et al., 25; Freymueller et al., 27] including 7 m of slip on the northernmost (Andaman) segment. We modeled the viscoelastic response for Earth models with a 4 8 km range of thicknesses for the elastic lithosphere, upper mantle (<67 km depth) Maxwell viscosity of η UM =5 1 16 1 21 Pa s, and lower mantle viscosity 1 21 Pa s. Figure 3 compares the GPS observations to the bestfitting Earth model, which assumed a 7 km elastic layer over a 5 1 17 Pa s upper mantle viscosity. Total displacements for the 24.98 27. epochs fit the GPS observations within uncertainties at some sites (e.g., ), but predicted displacements change by <4 cm across the Andaman network aperture. The GPS observations by contrast exhibit large displacement gradients. Modeled displacement directions and gradients are insensitive to assumed upper mantle viscosity, but magnitude is very sensitive, and lower viscosities yield correspondingly larger motions. Even where modeled and observed displacement vectors are similar, however, the time dependence of displacement is not. The decay time for the best-fitting η UM =5 1 17 model is 2.7+.8/.5 years at 95% confidence, after fitting an exponential to the first two years of modeled displacement at (Figure 3-inset a). This differs from the τ=.82 year East Up Vertical (m) 15 - -15 - -15 15 - -15 North15.4.35 a 25 26 27 Year Observed GPS 2 cm b.3.25.2.15.1 η UM =5x1 17 Pa s Modeled GPS 2 cm.3.4.5.6 Horizontal (m) Figure 3. GPS-observed and viscoelastic-modeled postseismic displacements. Red vectors/black bars are 24.98 27. horizontal/vertical displacements, respectively, from exponential fit of the GPS data, with scaled 95% confidence ellipses. Best-fit model of viscoelastic relaxation is depicted as blue/dark grey vectors. Insets: (a) Comparison of GPS coordinates (red circles) to best-fit model (blue line). (b) Vertical versus horizontal transient displacement; red is GPS estimate (with 2σ error), blue is viscoelastic model. Predictions for South Andaman sites are circled. timescale of the GPS data at >>99% confidence. Introducing a factor-of-3.3 lower viscosity to match the model decay timescale to the GPS observations would increase correspondingly the displacement vector magnitudes. 3.3. Postseismic slip To compare the data to a postseismic slip end-member, we modified slip modeling software developed for slow slip events in southern Mexico [Lowry, 26]. GPS positions x(t) were modeled as x(t) = x + V t + S ( ) ( ) ζ, t G x, ζ dζ. (2) Here, ζ denotes location on the fault surface, G is the deformation Green s function [Okada, 1985], and S describes the transient slip. Ideally S would be parameterized using frictional constitutive laws, but as a first approximation we let ( ) S( ζ, t) = S ( ζ) T t exp (3) τ in which S is the total transient slip anticipated following the earthquake. East Up 15 - -15 - -15 15 - -15 a 25 26 27 Year North15.35.3.25.2 Observed GPS 2 cm.15.1 Vertical (m).4 (poroelastic prediction) Modeled GPS 2 cm b.3.4.5.6 Horizontal (m) Slip Vector 1 2 3 Slip (m) 1 m Figure 4. Andaman data modeled as slip on the subduction thrust. Red vectors are 24.98 27. displacements from exponential fit of the GPS data, with scaled 95% confidence ellipses. Blue vectors are the best-fit model of postseismic slip. Red patches with thin black vectors indicate the magnitude and direction of modeled slip. Insets: (a) Comparison of GPS coordinates (red circles) to best-fit model (blue line). (b) Vertical versus horizontal transient displacement; red is GPS estimate (with 2σ error), blue is slip model. Dashed line schematically shows negative slope of poroelastic response.

X - 4 PAUL ET AL.: ANDAMAN POSTSEISMIC DEFORMATION We used the transient displacements A i estimated at each of the GPS sites i to invert for anomalous slip S on a 3 km mesh representation of a fault surface approximated from aftershocks and other well-located earthquakes on the plate interface. We inverted for both strike-slip and dip-slip. To regularize the solution, we required strike-slip not to exceed dip-slip and imposed a minimization of the total moment release. The latter constraint penalizes slip far from the GPS data constraint. The best-fit model of slip on the subduction interface is depicted in Figure 4. 4. Discussion and Conclusions 4.1. Relative Importance of Relaxation Processes Despite similarities, models of the three relaxation processes do differ. Viscoelastic models poorly match the directions, magnitudes, displacement gradients, and timedependence of the GPS measurements. A better fit to the observations might be obtained by further varying fault dislocation and viscosity parameterizations, but a fit as good as the fault slip model is unlikely to be physically viable. To match the timescale parameter τ =.82 years estimated for exponential decay would require a uniform Maxwell viscosity η 7 1 16 Pa s. Such a low viscosity is conceivable if the mantle is extremely water-rich and strain rate ɛ exceeds background rates by several orders of magnitude, given that η ɛ 1 n n for power-law creep. Matching other observations to a viscoelastic model would stretch limits of credibility even further. For example, sites,, and on South Andaman are spaced only 1 25 km apart, but their vertical displacements vary by up to 4%. The spatial wavelength of vertical response to a deformation point source is roughly equal to the depth, so large variations can be modeled as fault slip at the 35 4 km depth of the plate interface (Figure 4). Spatial wavelengths of viscoelastic response are filtered twice however: once to propagate stress from the dislocation to the depth of ductile creep, and again thence to the surface. The temperature structure for subduction of 1 Myr-old oceanic lithosphere would preclude creep above 8 km depth in this region, and the 7 km elastic lithosphere assumed in our best-fit model predicts only 7% variation in the vertical response of South Andaman sites (circled in Figure 3-inset b). These observations lead us to conclude viscoelastic relaxation does not dominate the first two years of Andaman near-field deformation. The strongest evidence against poroelasticity derives from displacement directions. Models of postseismic fluid diffusion [e.g., Fialko, 25] predict maximum uplift over loci of positive coseismic dilatational strain, and subsidence over contractional strain. Horizontal motions are away from the uplift and toward subsidence, vanishing where vertical displacements are largest and maximal between a fluid source and sink where vertical displacement is. Hence, on a plot of vertical versus horizontal displacement, the distribution of measurements should have negative slope. Inset b of Figures 3 4 exhibits a positive slope distribution. Although fault slip apparently dominates the Andaman Islands postseismic response, both viscoelastic and poroelastic processes must contribute. Viscoelastic modeling using reasonable viscosity structures predict displacements of up to 8 cm during the 24.98 27. epoch. Models of poroelastic response to earthquakes elsewhere also predict displacements of order several cm [e.g., Fialko, 25]. Hence, one reasonably might expect up to 25% error modeling the deformation as pure fault slip. Inspection of residuals at IGS sites used in our analysis suggest an additional 3% error in realization of the reference frame, related to largescale postseismic deformation. Separating these effects will require additional data that sample the differing temporal and spatial scales of the three processes, and future efforts will address this. 4.2. Implications of the Slip Model We modeled ten coseismic GPS vectors summarized by Freymueller et al. [27] using a minimum-moment constraint. The best-fit estimate of slip averaged 9.8 m within a rupture zone averaging about 9 km width (Figure 1-inset). The corresponding coseismic moment release for the 5 km Andaman segment is equivalent M w=8.6, comparable to the Freymueller et al. [27] M w=8.5 estimate in five rectangular dislocation patches, and also consistent with the M w=8.6 estimate from seismic data after combining the rapid (1 2 m in the first ten minutes) and slow (5 m over the next hour) slip averaged over a 16 km width [Lay et al., 25]. Despite possible contamination by other processes, the postseismic slip model is intriguing. The minimum-moment constraint images slip only where data are available, but there the slip magnitude and relative contribution of strike slip increase with depth. Slip vectors at depth mirror the surface GPS displacements, which are smaller and more trench-normal at western sites (Figure 4). Postseismic moment release is equivalent to a M w 7.5 earthquake, or about 1% of coseismic, and is probably larger given that a small fraction of the Andaman segment is sampled by data presented here. Postseismic slip overlaps slightly with the coseismic slip estimate (Figure 1-inset), but most of the moment release is further downdip. This pattern mirrors simulations of coseismic and postseismic slip in elastodynamic models of rateand state-dependent friction [Lapusta et al., 2, e.g.,], in which slip deficit accumulates in velocity-strengthening conditions during interseismic periods and catches up to coseismic slip over several years time. Andaman data prior to 24 are insufficient to assess interseismic coupling where we now identify postseismic slip. However modeling of interseismic slip and stress rates from GPS data at the Cocos- North America subduction boundary suggests locking above the frictional transition buffers stress accumulation in the shallow velocity-strengthening regime [Lowry, 26], thus enabling slip deficit to accumulate that can drive slip following an earthquake. Acknowledgments. We thank Professor V.K. Gaur (Indian Institute of Astrophysics), Mr. Samir Acharya (Society of Andaman and Nicobar Ecology) and Dr. T.V.R.S. Sharma (Central Agricultural Research Institute) for their partnership in this project. We also thank Dr. Anjan Battacharya for obtaining permission to set up GPS site, and Tom Herring for help and advice on GAMIT processing. This paper was improved by thoughtful suggestions of reviewers Cristophe Vigny and xxx. CERI publication #517. Support for this research was provided by NSF grants EAR-523319 and EAR-5359. References Fialko, Y., Evidence of fluid-filled upper crust from observations of postseismic deformation due to the 1992 M w7.3 Landers earthquake, J. Geophys. Res., 18 (B841), doi:1.129/24jb2,985, 25. Freed, A. M., R. Bürgmann, E. Calais, J. Freymueller, and S. Hreinsdottir, Implications of deformation following the 22 Denali, Alaska, earthquake for postseismic relaxation processes and lithospheric rheology, J. Geophys. Res., 111 (B141), doi:1.129/25jb3,894, 26. Freymueller, J. T., A. Shuler, C. P. Rajendran, A. Earnest, and K. Rajendran, Coseismic and preseismic displacements in the Andaman and Nicobar Islands, and implications for the regional tectonics, Bull. Seismol. Soc. Am., p. in press, 27.

PAUL ET AL.: ANDAMAN POSTSEISMIC DEFORMATION X - 5 Lapusta, N., J. R. Rice, Y. Ben-Zion, and G. Zheng, Elastodynamic analysis for slow tectonic loading with spontaneous rupture episodes on faults with rate- and state-dependent friction, J. Geophys. Res., 15, 23,765 23,789, 2. Lay, T., et al., The great Sumatra-Andaman earthquake of 26 December 24, Science, 38 (5725), 1127 1133, 25. Lowry, A. R., Resonant slow fault slip in subduction zones forced by climatic load stress, Nature, 442 (714), 82 85, 26. Meltzner, A. J., K. Sieh, M. Abrams, D. C. Agnew, K. W. Hudnut, J. P. Avouac, and D. H. Natawidjaja, Uplift and subsidence associated with the great Aceh Andaman earthquake of 24, J. Geophys. Res., 111, doi:1.129/25jb3,891, 26. Okada, Y., Surface deformation due to shear and tensile faults in a half-space, Bull. Seismol. Soc. Am.,, 1135 1154, 1985. Paul, J., et al., The motion and active deformation of India, Geophys. Res. Lett., 28 (4), 647 651, 21. Peltzer, G., P. Rosen, F. Rogez, and K. Hudnut, Poroelastic rebound along the Landers 1992 earthquake surface rupture, J. Geophys. Res., 13, 3,131 3,145, 1998. Pollitz, F. F., Gravitational-viscoelastic postseismic relaxation on a layered spherical Earth, J. Geophys. Res., 12, 17,921 17,941, 1997. Rundle, J. B., Viscoelastic crustal deformation by finite, quasistatic sources, J. Geophys. Res., 83, 5937 5945, 1978. Socquet, A., C. Vigny, N. Chamot-Rooke, W. Simons, C. Rangin, and B. Ambrosius, India and Sunda plates motion and deformation along their boundary in Myanmar determined by GPS, J. Geophys. Res., 111, doi:1.129/25jb3877, 26. Thatcher, W., and J. B. Rundle, A model for the earthquake cycle in underthrust zones, J. Geophys. Res., 84, 554 5556, 1979. Tse, S. T., and J. R. Rice, Crustal earthquake instability in relationship to the depth variation of frictional slip properties, J. Geophys. Res., 91, 9452 9472, 1986. J. Paul, Centre for Earthquake Research and Information, 3876 Central Ave, Memphis, TN 38152 USA. (jpuchkyl@memphis.edu) A. R. Lowry, Department of Geology, Utah State University, Logan, UT 84322-455, USA. (arlowry@cc.usu.edu) R. Bilham, Department of Geological Sciences, University of Colorado, Boulder, CO 839-399, USA. (Roger.Bilham@colorado.edu) S. Sen, Society for Andaman and Nicobar Ecology (SANE), Port Blair 74411, India. R. Smalley, Jr., Centre for Earthquake Research and Information, 3876 Central Ave, Memphis, TN 38152 USA.