role of initial stress rotations in rupture dynamics and ground motion

15
Click Here for Full Article Role of initial stress rotations in rupture dynamics and ground motion: A case study with implications for the Wenchuan earthquake Benchun Duan 1 Received 2 July 2009; revised 5 November 2009; accepted 2 December 2009; published 4 May 2010. [1] Motivated by observations in the 2008 M w 7.9 Wenchuan earthquake, we study effects of systematic changes in the principal stress orientation along the fault strike on rupture dynamics and ground motion using a 3D finiteelement method. Based on Andersons theory of faulting, we set up the initial stress field with rotations in stress orientations along strike for a dynamic rupture model of a shallow dipping fault. We find that initial stress rotations along strike can cause dramatic changes in rupture speed and produce distinct patterns in slip distribution and peak ground motion. When a mismatch (unfavorable for faulting) between fault geometry and initial stress orientations is encountered, rupture can spontaneously stop. Some firstorder features in the Wenchuan event may be partially caused by rotations in the principal stress orientation along strike, such as the rupture arrest at the northeast end and two severe destruction zones in the observed seismic intensity distribution. These results may have important implications for assessing seismic hazards imposed by faults with changes in the initial stress field along strike worldwide. Citation: Duan, B. (2010), Role of initial stress rotations in rupture dynamics and ground motion: A case study with implications for the Wenchuan earthquake, J. Geophys. Res., 115, B05301, doi:10.1029/2009JB006750. 1. Introduction [2] The rupture process of the 12 May 2008 M w 7.9 Wenchuan (China) earthquake was complex. Various kine- matic inversion studies [e.g., Ji et al., 2008; Wang et al., 2008; Zhang et al., 2009; N. Nishimura and Y. Yagi, un- published manuscript, 2008 (http://www.geol.tsukuba.ac.jp/ nisimura/20080512/)] demonstrate that systematic changes in faulting style along strike occurred in this event. For example, Zhang et al. [2009] inverted 48 worldwide stations of longperiod waveform data and found that the rupture started as thrust faulting in the southwest portion, changed to oblique faulting with comparable thrust and rightlateral strikeslip components as it progressed northeastward, and became predominately strikeslip faulting at the northeast end. Field investigations reveal that surface rupture from the event also exhibits the above feature of faulting style changes along strike [Xu et al., 2009; LiuZeng et al., 2009]. [3] Complex fault geometry was suggested as the cause for the 2002 Denali fault rupture to transfer from oblique faulting on the Susitna Glacier fault (NE trending) to pre- dominantly strikeslip motion on the Denali fault (EW trending) [Aagaard et al., 2004]. In that study, a uniform initial stress field is resolved onto differently oriented fault segments, resulting in changes in faulting style. However, the orientation of the Beichuan fault in the Wenchuan earthquake is essentially northeast. There are not signifi- cantly long fault segments departing from this trend. Step- overs/bends along strike [Xu et al., 2009] may cause local changes in faulting style that are present in some kinematic source models, but they cannot result in the above largescale, systematic changes. Furthermore, a single, planar dipping fault was used in some kinematic inversions [Zhang et al., 2009; Nishimura and Yagi, unpublished manuscript, 2008], while these kinematic models exhibit the above changes in faulting style. Therefore, complex fault geometry may not be the primary cause for the largescale, systematic faulting style changes in the Wenchuan earthquake. [4] Rotations of the principal stress field along strike prior to the Wenchuan earthquake likely play dominant roles in the faulting style changes. Changes in the orientation of the principal stress field along the Beichuan fault have been inferred from a shear wave splitting analysis [Liu et al., 2008] and a study on active fault slickensides [Du et al., 2009a]. Liu et al. [2008] analyzed seismic data recorded by a movable seismic array in the western Sichuan that was deployed from October 2006 to June 2008. They found that within the Longmen Shan fault zone, the fastwave direction is perpendicular to the faults in the south part, and it be- comes parallel to the faults in the north part. This finding strongly suggests that the maximum horizontal compressive stress (s 1 ) makes a very high angle to the fault trace (favoring thrust faulting) in the south portion of the Wenchuan rupture, the angle becomes shallower to the north (favoring oblique faulting), and it becomes very small (favoring predominantly strikeslip faulting) at the end of the rupture. Du et al. [2009a] analyzed measurements of 311 active fault slickensides and found a similar trend of s 1 rotations along the Beichuan fault. Rotations of the principal stress field along strike may be a result of com- 1 Center for Tectonophysics, Department of Geology and Geophysics, Texas A&M University, College Station, Texas, USA. Copyright 2010 by the American Geophysical Union. 01480227/10/2009JB006750 JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 115, B05301, doi:10.1029/2009JB006750, 2010 B05301 1 of 15

Upload: others

Post on 11-Feb-2022

3 views

Category:

Documents


0 download

TRANSCRIPT

ClickHere

for

FullArticle

Role of initial stress rotations in rupture dynamics and groundmotion:A case studywith implications for theWenchuan earthquake

Benchun Duan1

Received 2 July 2009; revised 5 November 2009; accepted 2 December 2009; published 4 May 2010.

[1] Motivated by observations in the 2008Mw 7.9 Wenchuan earthquake, we study effectsof systematic changes in the principal stress orientation along the fault strike on rupturedynamics and ground motion using a 3‐D finite‐element method. Based on Anderson’stheory of faulting, we set up the initial stress field with rotations in stress orientations alongstrike for a dynamic rupture model of a shallow dipping fault. We find that initial stressrotations along strike can cause dramatic changes in rupture speed and produce distinctpatterns in slip distribution and peak ground motion. When a mismatch (unfavorable forfaulting) between fault geometry and initial stress orientations is encountered, rupture canspontaneously stop. Some first‐order features in the Wenchuan event may be partiallycaused by rotations in the principal stress orientation along strike, such as the rupture arrestat the northeast end and two severe destruction zones in the observed seismic intensitydistribution. These results may have important implications for assessing seismic hazardsimposed by faults with changes in the initial stress field along strike worldwide.

Citation: Duan, B. (2010), Role of initial stress rotations in rupture dynamics and ground motion: A case study withimplications for the Wenchuan earthquake, J. Geophys. Res., 115, B05301, doi:10.1029/2009JB006750.

1. Introduction

[2] The rupture process of the 12 May 2008 Mw 7.9Wenchuan (China) earthquake was complex. Various kine-matic inversion studies [e.g., Ji et al., 2008; Wang et al.,2008; Zhang et al., 2009; N. Nishimura and Y. Yagi, un-published manuscript, 2008 (http://www.geol.tsukuba.ac.jp/∼nisimura/20080512/)] demonstrate that systematic changesin faulting style along strike occurred in this event. Forexample, Zhang et al. [2009] inverted 48 worldwide stationsof long‐period waveform data and found that the rupturestarted as thrust faulting in the southwest portion, changedto oblique faulting with comparable thrust and right‐lateralstrike‐slip components as it progressed northeastward, andbecame predominately strike‐slip faulting at the northeastend. Field investigations reveal that surface rupture from theevent also exhibits the above feature of faulting stylechanges along strike [Xu et al., 2009; Liu‐Zeng et al., 2009].[3] Complex fault geometry was suggested as the cause

for the 2002 Denali fault rupture to transfer from obliquefaulting on the Susitna Glacier fault (NE trending) to pre-dominantly strike‐slip motion on the Denali fault (EWtrending) [Aagaard et al., 2004]. In that study, a uniforminitial stress field is resolved onto differently oriented faultsegments, resulting in changes in faulting style. However,the orientation of the Beichuan fault in the Wenchuanearthquake is essentially northeast. There are not signifi-cantly long fault segments departing from this trend. Step-

overs/bends along strike [Xu et al., 2009] may cause localchanges in faulting style that are present in some kinematicsource models, but they cannot result in the above large‐scale, systematic changes. Furthermore, a single, planardipping fault was used in some kinematic inversions [Zhanget al., 2009; Nishimura and Yagi, unpublished manuscript,2008], while these kinematic models exhibit the abovechanges in faulting style. Therefore, complex fault geometrymay not be the primary cause for the large‐scale, systematicfaulting style changes in the Wenchuan earthquake.[4] Rotations of the principal stress field along strike prior

to the Wenchuan earthquake likely play dominant roles inthe faulting style changes. Changes in the orientation of theprincipal stress field along the Beichuan fault have beeninferred from a shear wave splitting analysis [Liu et al.,2008] and a study on active fault slickensides [Du et al.,2009a]. Liu et al. [2008] analyzed seismic data recordedby a movable seismic array in the western Sichuan that wasdeployed from October 2006 to June 2008. They found thatwithin the Longmen Shan fault zone, the fast‐wave directionis perpendicular to the faults in the south part, and it be-comes parallel to the faults in the north part. This findingstrongly suggests that the maximum horizontal compressivestress (s1) makes a very high angle to the fault trace(favoring thrust faulting) in the south portion of theWenchuan rupture, the angle becomes shallower to thenorth (favoring oblique faulting), and it becomes very small(favoring predominantly strike‐slip faulting) at the end ofthe rupture. Du et al. [2009a] analyzed measurements of311 active fault slickensides and found a similar trend ofs1 rotations along the Beichuan fault. Rotations of theprincipal stress field along strike may be a result of com-

1Center for Tectonophysics, Department of Geology and Geophysics,Texas A&M University, College Station, Texas, USA.

Copyright 2010 by the American Geophysical Union.0148‐0227/10/2009JB006750

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 115, B05301, doi:10.1029/2009JB006750, 2010

B05301 1 of 15

plex material movements along the east edge of the Tibetplateau.[5] Initial stress field, fault geometry, and friction laws are

three major ingredients of a dynamic source model [Harriset al., 2009]. Effects of complex fault geometry on rupturedynamics and near‐field ground motion have been a focusof recent studies [e.g., Harris and Day, 1993; Aochi et al.,2000; Kame et al., 2003; Duan and Oglesby, 2005, 2006,2007]. Uniform regional stress field, which is resolved ontodifferently oriented fault segments, is usually assumed inthese studies. The role of systematic rotations of the prin-cipal stress field along strike in rupture dynamics has notbeen explored in the literature, although effects of stochasticfault stress have been examined [Oglesby and Day, 2002].This study attempts to fill this gap by a case study motivatedby the observations in the Wenchuan earthquake. Under-standing roles of stress rotations along strike in rupturepropagation and ground motion may have importantimplications for assessing seismic hazards imposed by faultswith stress rotations along strike worldwide. We build ourtarget model on the basis of some observations in theWenchuan earthquake, specifically the stress rotations alongstrike and the shallow dipping fault geometry at depth.However, we do not attempt to build a comprehensivedynamic source model for the earthquake, which is beyondthe scope of this study.

2. Model

[6] Figure 1 shows the fault geometry, the coordinatesystem, and the orientation of the maximum horizontalcompressive stress s1 in the target model OBLIQUE. Aplanar dipping fault with a dip angle b = 33° (e.g., Nishimura

and Yagi, unpublished manuscript, 2008) is embedded in auniform 3‐D half‐space with a flat surface. The fault is300 km long along strike and 40 km wide downdip. Thefault strike coincides with the x axis of the coordinatesystem. Although this fault geometry is a significantlysimplified representation of the actual complex fault ge-ometry in the Wenchuan earthquake, it captures the featureof shallow dipping angle at depth and fits teleseismic datawell (e.g., Nishimura and Yagi, unpublished manuscript,2008). On the basis of current understanding of stressdistribution in the upper crust, we set up initial stresses onthe fault plane, including normal, strike‐slip shear, and dip‐slip shear stresses, in our target model OBLIQUE. Severalsimple principles are adopted. First, we start from theprincipal stress components of the initial stress field. Rela-tive sizes and orientations of three principal stress compo-nents are chosen based on Anderson’s theory of faulting[Anderson, 1951] to produce the desired faulting style onfault segments. Second, we take into account the depthdependence of the stress field and pore fluid effects. Thus,the magnitude of effective principal stress components in-creases with depth. Third, initial fault stresses at a faultpoint are obtained by resolving the above effective principalstress field onto the fault plane.[7] As a simple case and a first‐order approximation to the

changes in the stress orientation observed in the Wenchuanearthquake, we assume that the principal stress field rotatesat x = 50 km and x = −100 km along the strike (Figure 1),separating the 300 km long fault into the SW 100 km long,middle 150 km long, and NE 50 km long segments(denoted, respectively, SWS, MDS, and NES). The orien-tation of the principal stress field is characterized by theangle a between the maximum horizontal compressive stress

Figure 1. Fault geometry, the coordinate system, and the most compressive horizontal stress (s1) orien-tations in the target model OBLIQUE in this study. a is the angle between s1 and the fault trace and is90°, 60°, and 30° on three segments of the fault, respectively. The dip angle of the fault is b = 33°. Thefault is 300 km long along strike and 40 km wide along the dip direction. The star is hypocenter in sim-ulated ruptures, which is at an along‐strike distance of 125 km and a vertical depth of 10 km.

DUAN: STRESS ROTATIONS IN THE WENCHUAN EARTHQUAKE B05301B05301

2 of 15

s1 and the fault trace. The angle a is chosen to be 90°, 60°,and 30°, respectively, for the above three segments, resultingin pure‐thrust, oblique (thrust and strike‐slip), and pre-dominantly strike‐slip faulting on SWS, MDS, and NES,respectively.[8] The principal stress components at a fault point are

calculated as follows [e.g., Ranalli, 1995]. First, the verticalprincipal stress component is assumed to be lithostatic:

�3 ¼ �gz; ð1Þ

where r, g, and z are density, gravity, and depth, respec-tively. Second, a stress ratio R is defined by a nominalfrictional coefficient m0 that characterizes the initial stressstate on the fault plane:

R ¼ �max=�min ¼ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1þ �2

0

� �q� �0

� ��2

; ð2Þ

where smax and smin are the maximum and minimumprincipal stresses, respectively. Taking compressive stress aspositive, we will always have smax = s1 in this study, whilesmin can be s2 or s3, depending on faulting style. Third,choices of s2 are made so that all three principal stresscomponents can be determined. The choice of s2 is some-what arbitrary. We choose s2 = s3 for pure thrust fault onSWS and oblique MDS faulting on MDS, and s2 = 2

1þRs3for predominantly strike‐slip faulting on NES. The coeffi-cient 2/(1 + R) here comes from the fact that the verticalprincipal stress s3 is the intermediate stress for strike‐slipfaulting, and it is equal to s2 + d(s1 − s2) with a specialvalue of the coefficient d = 1/2. The above choice ofs2 results in s1 = Rs3 on segments of SWS and MDS ands1 = 2R

1þRs3 on NES. We remark that at a given depth alongthe fault strike, the magnitudes of s1 and s2 on NES aredifferent from those on SWS and MDS in the model,although the magnitude of s3 is same. This is a consequenceof different roles of the vertical principal stress s3 in differentfaulting styles [Anderson, 1951], while s3 is the basis indetermining the magnitude of s1 and s2 as above.

Figure 2. Stresses (MPa) on the fault plane in two models: (left) OBLIQUE and (right) THRUST. sn, ts,td, and Dt are initial normal (magnitude), initial strike‐slip (right‐lateral) shear, initial dip‐slip (thrust)shear stresses, and shear stress change (final minus initial), respectively. Along‐strike profiles are atdowndip depth of 20 km, and downdip sections are at along‐strike distance of 0 km. Initial yield stressis the product of initial normal stress and static frictional coefficient. Three stress regimes along strikeexist in OBLIQUE. The star denotes the hypocenter.

DUAN: STRESS ROTATIONS IN THE WENCHUAN EARTHQUAKE B05301B05301

3 of 15

[9] Taking into account pore fluid pressure pf = ls3 =lrgz, where l is the ratio between fluid pressure andlithostatic stress, we obtain effective principal stresses as

�e3 ¼ �3 � pf ¼ 1� �ð Þ�gz for SWS; MDS; NES;

�e2 ¼ �e

3; �e1 ¼ R�e

3 for SWS and MDS;

�e2 ¼

2

1þ R�e3; �e

1 ¼2R

1þ R�e3 for NES;

ð3Þ

Fault stresses, including normal sn, strike‐slip shear ts, anddip‐slip shear td components, can be obtained by resolvingthe above effective principal stress field at a fault point ontothe fault plane. Notice that even without any changes in themagnitude of the three principal stress components at a

given depth from SWS to MDS, the magnitude of theresolved fault stress components changes at the junctionbetween SWS and MDS (Figure 2, left) due to the rotationof the two horizontal principal stress components.[10] In dry rock, l = 0; while l � 0.9 can be reached in

water‐saturated sediments. We choose l = 0.7 in this studyto produce slip amplitude comparable with that obtainedfrom kinematic inversions for the Wenchuan earthquake

Figure 3. A time‐weakening friction model in which fric-tional coefficient m drops from static value ms to dynamicvalue md over a critical time period T0.

Figure 5. A schematic mesh for the models in this study.Most elements are hexahedrons. Wedge elements are createdonly on the two sides of the fault plane to accommodate theshallow dipping fault geometry.

Figure 4. Split nodes of a dipping fault element surface. Alocal coordinate system with three axes coinciding withthree unit vectors of the fault surface (normal n, strike s,and dip d), rather than the global coordinate system (x, y, z),is used to implement the traction‐at‐split‐node approach inEQdyna. The fault is normally an idealized mathematicalplane for shear sliding. It is shown as two separated planesfor split‐node illustration purpose.

Table 1. Model Parameters in This Study

Parameter Value

Density 2670 kg/m3

P wave velocity 6000 m/sS wave velocity 3464 m/sFault dip b 33°Fault length along strike 300 kmFault downdip width 40 kmFault element size 500 mPore pressure ratio l 0.7Time weakening T0 0.72 sCritical radius rc 5 kmNucleation V0 2 km/sFriction for stress m0 0.5Static friction ms 0.6Dynamic friction md 0.43Damping parameter " 0.1Time step Dt 0.02 sTermination time 100 s

DUAN: STRESS ROTATIONS IN THE WENCHUAN EARTHQUAKE B05301B05301

4 of 15

(e.g., Nishimura and Yagi, unpublished manuscript, 2008).This results in the pore pressure in the model being abovehydrostatic. We choose m0 = 0.5, which gives R = 2.62 byequation (2). This value of R is used from the surface to thehypocenter depth (10 km vertically, or 18.36 km along thedowndip direction), below which R linearly decreases to 2.3

at the bottom of the fault to avoid dramatic increase in stressdrop on the deeper portion of the fault, which might resultin unrealistic large slip at the bottom of the fault.[11] The initial fault stresses in the target model

OBLIQUE obtained by the above procedure and parametersare shown in Figure 2 (left). To facilitate examination of

Figure 6. Contour plots of rupture time for (a) OBLIQUE and (b) THRUST. Rupture time at a faultnode is defined as the time when the amplitude of slip velocity first reaches �0.001 m/s.

Figure 7. Snapshots of (left) the thrust slip rate and (right) the strike‐slip rate for (a) OBLIQUE and(b) THRUST at 20, 40, 60, and 80 s. Rake rotation near the free surface (nonzero strike‐slip velocity)is obvious in THRUST. A gentle slope of the rupture front and a low slip velocity zone near the surface atlater times can be observed in OBLIQUE.

DUAN: STRESS ROTATIONS IN THE WENCHUAN EARTHQUAKE B05301B05301

5 of 15

effects of the stress rotations, we also perform a dynamicsimulation on a model THRUST as a reference. In THRUST,the initial stress field along the strike does not rotate andthe principal stress field of SWS in OBLIQUE is appliedalong the entire length of the fault (Figure 2, right). Nor-mal and shear fault stresses in both models increase withdepth, with the exception of zero strike‐slip shear stress inTHRUST on the entire fault. Notice that the only differ-ence between OBLIQUE and THRUST is in the initialstress field as discussed above, and all other parameters areexactly the same.[12] Two stress profiles are also shown in Figure 2. One is

along the horizontal central line (Figure 2, second frombottom), and the other is the vertical central line (Figure 2,bottom) on the fault plane. In Figure 2, shear stress is thevector magnitude of the two shear traction components (tsand td) on the fault, and yield stress is the product of normalstress and static friction coefficient ms. Given uniformlydistributed ms on the entire fault, yield stress is a proxy fornormal stress. It can be seen that strength excess, which isthe difference between initial yield and initial shear stressesat a fault point, is much larger on NES compared with thaton SWS and MDS in OBLIQUE. Final shear stress at thetermination of dynamic simulations is also displayed inFigure 2. Shear stress changes Dt (final minus initial) onthe entire fault plane due to dynamic rupture are also givenin Figure 2.

[13] A time‐weakening frictional law [Andrews, 2004],shown in Figure 3, is used in this study. In this frictionalmodel, the frictional coefficient drops from a static value msat the time shear stress reaches yield stress to a dynamicvalue md over a critical time period T0. This time‐weakeninglaw with a constant T0 is equivalent to widely used slip‐weakening frictional laws in which frictional coefficient is afunction of slip rather than time: if off‐fault material be-haves elastically and stress drop on a fault plane is constant,it is equivalent to a linear slip‐weakening law with thecritical slip distance D0 proportional to the square root ofrupture length [Andrews, 2004]; if off‐fault material is al-lowed to yield (i.e., elastoplastic behavior), it is equivalentto a linear slip‐weakening law with a constant D0 [Duan andDay, 2008]. Because we assume that off‐fault material be-haves elastically in this study, the time‐weakening law witha constant T0 can avoid continuous shrinking of the cohesivezone width associated with a slip‐weakening law with aconstant D0 [e.g., Day et al., 2005] or complexity in numer-ical calculations associated with a slip‐weakening law withvarying D0. A T0 value of 0.72 s is used in this study, whichis the time for an S wave to propagate five fault elements, toensure good resolution of the cohesive zone at the rupturefront [e.g., Day et al., 2005].[14] Homogeneous material with r = 2670 kg/m3, P wave

velocity Vp = 6000 m/s, and S wave velocity Vs = 3464 m/sis assumed in the entire model to isolate effects of the initial

Figure 8. Time‐space plots of slip rate (magnitude) at downdip depths of 1, 20, and 35 km for (left)OBLIQUE and (right) THRUST. The white line indicates the shear wave speed in the models. Supershearrupture along the middle segment can be observed at 1 and 20 km depths in OBLIQUE.

DUAN: STRESS ROTATIONS IN THE WENCHUAN EARTHQUAKE B05301B05301

6 of 15

stress rotations and the dipping fault geometry from othersources of complexity. These parameters and other com-putational parameters are summarized in Table 1.

3. Method

[15] We use an explicit finite‐element method EQdyna[Duan andOglesby, 2006, 2007;Duan andDay, 2008;Duan,2008a, 2008b; Harris et al., 2009] to simulate spontaneousrupture propagation on faults and wave propagation in themedium. The formulation of EQdyna follows standard finite‐element procedures [e.g., Hughes, 2000] with the addition ofa treatment of faulting boundary condition. A general de-scription of the formulation is given by Duan and Oglesby[2006]. Dalguer and Day [2006] showed that the accuracyof spontaneous rupture solutions largely depends on how thefaulting boundary is treated. They found that the traction‐at‐split‐node (TSN) method works best among three schemesthey examined. Early versions of EQdyna [Duan andOglesby, 2006, 2007] followed the implementation of TSNsummarized by Andrews [1999] in which fault slip and sliprate are explicitly calculated. However, this implementationrequires separate treatments for initiation of sliding, contin-uous sliding, and arrest of sliding. A new formulation of TSNproposed byDay et al. [2005] treats fault behavior at all timesin a concise and robust framework. New versions of EQdyna[Duan, 2008a, 2008b; Duan and Day, 2008] follow thisimplementation of TSN. The most updated version ofEQdyna used in this study adopts the OpenMP parallelprogramming technique to make use of modern chip multi-processor systems [Wuet al., 2009]. The code has been verifiedin the Southern California Earthquake Center dynamic code

validation exercise on a set of benchmark problems [Harriset al., 2009]. The most recent benchmark problems examinedynamic ruptures on a dipping (60°) normal fault withdepth‐dependent stress distribution on the fault.[16] The implementation of TSN in the current version of

EQdyna essentially follows that of Day et al. [2005]. Incontrast to the formula given in a global coordinate system(x, y, z) by Day et al. [2005], we solve for the trial tractionbetween a pair of split nodes in a local coordinate system(n, s, d) defined by the unit normal, strike, and dip vectorsof the fault plane (Figure 4):

~Tv ¼�t�1MþM� _uþv � _u�v

� �þ�t�1 uþv � u�v� �

�vn� �þM�Rþ

v �MþR�v

a Mþ þM�ð Þþ T0

v ; v ¼ s; d; n; ð4Þwhere dvn is the Dirac delta function, M, Rv, _uv, and uvwith superscript plus and minus signs are mass, elasticrestoring force, particle velocity, and particle displacementof plus‐side and minus‐side split nodes, respectively, Tv

0 isthe initial stress (in equilibrium), a is the area of the faultsurface associated with the split node pair in consideration,and Dt is dynamic simulation time step. Similarly, the truetraction is also defined in the local coordinate system as

Tv ¼

~Tv; v ¼ s; d; T � �c

�c~Tv=T ; v ¼ s; d; T > �c

~Tv; v ¼ n; ~Tn � 0

0; v ¼ n; ~Tn > 0;

8>>>>>>>><>>>>>>>>:

ð5Þ

Figure 9. Slip rate profiles along strike at downdip depths of 1, 10, 20, and 35 km at 70 s in OBLIQUE.Two rupture fronts (supershear and Rayleigh) at shallow depth result in a low slip rate zone between thetwo fronts.

DUAN: STRESS ROTATIONS IN THE WENCHUAN EARTHQUAKE B05301B05301

7 of 15

where T = [(~T s)2 + (~Td)

2]1/2 is the magnitude of shearstress on the fault plane, and tc is nonnegative frictionalstrength. Here, we consider the compressive normal stressas negative.[17] For computational efficiency, we primarily use

hexahedron elements in EQdyna. To accommodate com-plexity in fault geometry and velocity structure, we use thedegeneration technique [e.g., Hughes, 2000] to obtainwedge elements and/or tetrahedron elements whenevernecessary. For example, to accommodate the shallow dip-ping fault geometry in this study, we adopt wedge elementson the two sides of the fault plane. After degeneration, onehexahedron element is split into two wedge elements asillustrated in Figure 4.[18] A schematic version of the mesh used in this study is

shown in Figure 5, in which wedge elements only exist onthe two sides of the fault plane. In this mesh, we introduce abuffer region surrounding the fault region to prevent re-flections from the model edges (except the free surface)from propagating back to a space‐time subdomain of in-terest. Split‐node pairs along the three edges (except the toptrace) of the 300 km × 40 km fault plane are pinned togetherby a high static friction coefficient ms (e.g., 1000). A nom-inal element size of 500 m, which is the grid spacing alongthe x (strike) and downdip directions of the fault plane, is

used. This results in a grid spacing of 419.3 m along the ydirection and a grid spacing of 272.3m along the z direction inthe fault region. If we take 10 grid intervals (i.e., 11 nodes) pershear wavelength, the highest frequency in seismic wavesaccurately simulated in this study is about 1.3 Hz. Theelement number in the models is about 47 million. Witha dynamic simulation time step of 0.02 s and a terminationtime of 100 s, one dynamic simulation takes about 25 h and42 GB memory to run on a SUN server with four dual‐coreAMD Opteron processors.[19] In EQdyna, there are two mechanisms to initiate

rupture, and both utilize the concept of a critical crack. Oneis to assign the initial shear stress within this nucleationpatch to be higher than the yield stress. The other is to assigna fixed rupture velocity V0 within this patch (i.e., within acircular patch with a radius of rc) for expansion of therupture front, which is adopted in this study. Outside of thisnucleation patch, the rupture propagates spontaneously.[20] Spontaneous rupture propagations are nonlinear pro-

blems. Since high‐frequency components of the solution donot simply superimpose linearly on low‐frequency compo-nents, the inevitable numerical inaccuracies at wavelengthsnear the resolution limit of the mesh, if not controlled, canpotentially contaminate the solution at all space‐time scales.We use stiffness‐proportional Rayleigh damping [e.g.,

Figure 10. Slip distribution on the entire fault plane for (a) THRUST and (b) OBLIQUE and surface slipfor (c) OBLIQUE and (d) THRUST. Arrows in Figures 10a and 10b denote the slip direction of hangingwall relative to footwall. Slip rake rotation near the surface results in surface strike‐slip rupture even forpure thrust faulting (in the sense of zero initial strike‐slip shear stress).

DUAN: STRESS ROTATIONS IN THE WENCHUAN EARTHQUAKE B05301B05301

8 of 15

Hughes, 2000; Duan and Oglesby, 2006; Duan and Day,2008] to selectively damp wavelengths near the resolutionlimit of the numerical mesh. The damping coefficient q canbe specified through a nondimensional parameter ", suchthat q = "Dt. See Table 1 for a summary of parameters usedin this study.

4. Results

[21] We examine both dynamic rupture propagation andground motion from the target model OBLIQUE and thereference model THRUST. We consider these models asgeneric faulting models in this section and discuss theimplications of results, especially for the 2008 Mw 7.9Wenchuan earthquake, in the next section.

4.1. Dynamic Rupture Propagation

4.1.1. Rupture Speed and Slip Rate[22] Effects of stress rotations on dynamic rupture prop-

agation and fault slip rate are demonstrated by rupture timecontours and slip rate snapshots on the fault plane in Figures 6and 7, respectively, and by time‐space distributions of the sliprate magnitude and slip rate profiles at different downdipdepths in Figures 8 and 9, respectively. First, the stressrotation at the junction between SWS and MDS (x = 50 km)

in OBLIQUE causes a dramatic change in rupture speed.The rupture accelerates near the free surface, while it de-celerates at the bottom of the fault at this junction. This canbe clearly seen by comparing rupture time contours betweenOBLIQUE and THRUST around the junction in Figure 6.The change in rupture speed at the junction at differentdepths is further illustrated by the time‐space plot of sliprate in Figure 8. This results in a profound change in theshape of the rupture front (Figures 6 and 7). The rupturefront is relatively steep on SWS and becomes shallower onMDS in OBLIQUE.[23] Second, the stress rotation at the above junction

triggers a supershear rupture transition at shallow depth inOBLIQUE. Although rupture near the free surface is fasterthan that at depth on SWS (Figure 6) in both OBLIQUE andTHRUST, it is still subshear on SWS, and this feature holdsalong the entire length of the fault in THRUST. However,the rupture in OBLIQUE jumps to supershear at shallowdepth when it passes through the junction, as shown in thetime‐space plot of slip rate at 1 km downdip depth inFigure 8 (top left). A manifestation of the supershear ruptureafter the junction is the development of two rupture fronts inOBLIQUE, as labeled in Figure 9. One rupture front pro-pagates at a speed above the shear wave speed (the super-shear front) and the other at the Rayleigh wave speed (the

Figure 11. Snapshots of (left) horizontal and (right) vertical ground velocity at 20, 40, 60, and 80 s for(a) OBLIQUE and (b) THRUST. The black line is the fault trace. Circle, triangular, cross, and plus sym-bols are epicenter, Chengdu, Wenchuan, and Beichuan cities if the model is applied to the 2008 Wench-uan earthquake.

DUAN: STRESS ROTATIONS IN THE WENCHUAN EARTHQUAKE B05301B05301

9 of 15

Rayleigh front). Slip rate between the two fronts is relativelylow, as shown in Figures 7–9, from which we can see thatsupershear rupture occurs at downdip depth up to ∼20 km.The supershear front becomes stronger (i.e., the peak of sliprate increases) with depth in this depth range, while theRayleigh front has no obvious change in strength with depth(Figure 9). High‐frequency noises seen in the profiles of the10 and 20 km depths in Figure 9 may be associated with thestrong supershear rupture front over the intermediate depth,which could generate a wide range of strong high‐frequencyseismic radiations. The finite size of numerical elements willresult in some numerical noise even with the presence of thenumerical damping mechanism in EQdyna.[24] Third, the stress rotation at the junction between

MDS and NES (x = −100 km) in OBLIQUE results in aspontaneous arrest of the rupture (Figure 6). The ruptureessentially dies out at the junction on deeper parts of thefault, although it penetrates into NES near the free surface.The spontaneous arrest of the rupture appears to be causedby the higher strength excess along NES (Figure 2). Thishigh strength excess results from the combination of thestress orientation and the shallow dipping fault geometry. Ifeither the dip of the fault along NES is steep or the anglebetween the maximum compressive stress and the fault traceis big, the rupture would not stop there. Notice that weterminate the simulations at 100 s, at which the rupture inTHRUST is still steadily propagating as it has not reachedthe northeast end of the fault.[25] A common feature in the slipping‐zone distribution

in OBLIQUE and THRUST from Figures 7–9 is that the

slipping area on the fault plane at a given time decreaseswith depth. This appears to be associated with the depth‐dependent stress distribution, which results in larger strengthexcesses on deeper parts of the fault (Figure 2, bottom).4.1.2. Slip Distribution and Stress Drop[26] Slip distribution on the fault plane is shown in

Figure 10. The slip direction (motion of hanging wall relativeto footwall) is displayed by arrows. The slip distribution onthe fault plane in THURST (Figure 10a) is typical for adynamic model with planar fault geometry and uniformstress distribution along strike: one slip patch develops,while the stress rotation between SWS andMDS at x = 50 kmin OBLIQUE results in two large slip patches with reducedslip in between (Figure 10b). It is clear that different prin-cipal stress orientations along SWS and MDS produce dif-ferent faulting styles on the two fault segments (arrows inFigures 10a and 10b). At the junction between MDS andNES, slip penetrates farther onto NES near the free surface,while it stops at depth. These features are also observed insurface rupture. In OBLIQUE (Figure 10c), two peaks existin magnitude of surface slip, and surface slip penetrates ontoNES for several tens of kilometers.[27] Profound slip rake rotation near the free surface can

be clearly seen in surface rupture of THRUST (Figure 10d)and along SWS of OBLIQUE (Figure 10c). A significantamount of the strike‐slip component (i.e., ∼2 m or larger inamplitude, right‐lateral positive) develops under the condi-tion of zero initial strike‐slip stress. This phenomenon hasbeen documented in previous studies [e.g., Spudich, 1992;Oglesby et al., 2000; Oglesby and Day, 2001] and is caused

Figure 12. Three components and magnitude of peak ground acceleration (PGA) from (left) OBLIQUEand (right) THRUST. See legend to Figure 11 for the symbols.

DUAN: STRESS ROTATIONS IN THE WENCHUAN EARTHQUAKE B05301B05301

10 of 15

by dynamic stresses that are in a different direction than thestatic stresses.[28] The above slip distributions in the two models

correlate well with the stress drop distributions shown inFigure 2. One large stress drop patch develops in THRUST,while two large stress drop patches develop in OBLIQUE,with reduced stress drop in between. Shear stress increasesdramatically at the northeast end of the rupture in OBLIQUE:a red zone with stress increase of 5 MPa or larger is verynarrow at deeper parts of the fault and expands laterallytoward the free surface. This zone is terminated below thesurface due to the slip penetration. Notice that a large shearstress increase zone at the NE end of the THRUST ruptureis associated with the propagating rupture.

4.2. Ground Motion

4.2.1. Wave Propagation[29] Snapshots of horizontal and vertical ground velocity

from OBLIQUE and THRUST are compared in Figure 11.Ground motion is exactly same at 20 s and is similar at 40 sbetween the two models. At both times, both strongesthorizontal and vertical ground shaking are associated withthe rupture front, which propagates at nearly Rayleigh wavespeed. Wavefields at 60 and 80 s in THRUST exhibit theseismic radiation field of a steadily propagating rupture at asubshear speed. On the other hand, wavefields at these timesin OBLIQUE show several prominent features that areassociated with changes in rupture speed at the junctionbetween SWS and MDS (x = 50 km). First, the strongest

horizontal ground shaking is associated with the supershearrupture front, while the strongest vertical ground shaking isassociated with the Rayleigh rupture front. Second, a ringlikezone in ground velocity centered at the junction is clearlyseen at 60 s and is also discernible at 80 s (Figure 11, arrow),which is absent in THRUST. This wave is emitted by thetransition from pure thrust to oblique faulting at x = 50 km,where the rupture speed changes dramatically. It is obviousthat large ground shaking is concentrated on the hangingwall side of the fault, which is an effect of dipping faultgeometry on near‐field ground motion [e.g., Oglesby et al.,2000].4.2.2. Peak Ground Motion[30] Low‐frequency (i.e., up to 1.3 Hz) peak ground

acceleration (PGA), peak ground velocity (PGV), and finalsurface displacement field are shown in Figures 12–14,respectively. The straight line in these figures denotes thefault trace. By comparing these quantities from OBLIQUEand THRUST, we find that two prominent features are as-sociated with the stress rotation at x = 50 km between SWSand MDS. First, two patches of large ground displacement(Figure 14) develop in OBLIQUE, particularly in the ver-tical component and the vector magnitude. This distributionin ground displacement correlates well with the slip distri-bution on the fault, and both are caused by the stress rotationat x = 50 km. Second, in addition to large PGA along the faulttrace, there is a zone (Figure 12, arrows) with relatively largePGA extending from the fault trace at the junction betweenSWS and MDS to the hanging wall side of the fault (negative

Figure 13. Three components and magnitude of PGV from (left) OBLIQUE and (right) THRUST. Seelegend to Figure 11 for the symbols.

DUAN: STRESS ROTATIONS IN THE WENCHUAN EARTHQUAKE B05301B05301

11 of 15

across‐fault distance) in OBLIQUE. This zone can be seen inx and z components and the vector magnitude. It appears thatthis feature is an effect of the stress rotation at x = 50 kmon PGA because this feature is absent in THRUST.Contrast in ground velocity between the hanging wall andfootwall can be observed in PGV (Figure 13). Once again,this is an effect of dipping fault geometry on near‐fieldground motion [e.g., Oglesby et al., 2000].4.2.3. Waveforms[31] Effects of the stress rotation between SWS and MDS

on waveforms can be seen in Figure 15, which shows timehistories of acceleration and velocity from OBLIQUE(dashed lines) and THRUST (solid lines) at four selectedstations. The four stations are along the profile of x = −25 km.Two of them are on the footwall with distances from the faulttrace of 1 and 10 km, respectively. The other two (negativevalues in y coordinate) are on the hanging wall and aresymmetric with the above two stations about the fault trace.The time range in Figure 15 is between 45 and 75 s. Twodistinct features in the waveforms of OBLIQUE are associ-ated with the stress rotation at x = 50 km. One is the enhancedhigh‐frequency content (more wiggles and/or kinks) in thewaveforms at these near‐field stations. These overall en-hanced high‐frequency ground motions are primarily causedby supershear rupture propagation on MDS due to the stressrotation. The other feature is that there are late arrivals inwaveforms. These late arrivals are high‐frequency radiationsfrom the junction between SWS and MDS on the fault andthey primarily appear in the acceleration waveforms. Larger

ground motion on the hanging wall (negative y coordinatevalues) and a slower rate of decay away from the faulttrace (y = 0 km) compared with those on the footwall areclearly seen in these waveforms in both models. This isprimarily the effect of the shallow dipping fault geometry[e.g., Oglesby et al., 2000].

5. Discussion

[32] The target model OBLIQUE is motivated by ob-servations of the 2008 Wenchuan earthquake. AlthoughOBLIQUE is a significantly simplified dynamic modelcompared with observations of the earthquake, it capturestwo important factors, namely dipping fault geometry andinitial stress rotations along strike, which appear to dominatesome first‐order features in rupture propagation and near‐field ground motion in the event. Here, we mainly examineimplications of the results from OBLIQUE for theWenchuanearthquake. We also discuss some general implications of theresults in a broader context.

5.1. A Possible Mechanism of Rupture Arrest at theNortheast End of the Wenchuan Rupture: MismatchBetween Initial Stress Orientation and Fault Geometry

[33] In our model OBLIQUE, the 300 km long and 40 kmwide fault is pinned at the two lateral boundaries and thebottom boundary by very high static friction (thus very highfrictional strength). Thus, the model does not provide anyinformation about why rupture stops at the southwest end of

Figure 14. Three components and magnitude of ground displacement field from (left) OBLIQUE and(right) THRUST. See legend to Figure 11 for the symbols.

DUAN: STRESS ROTATIONS IN THE WENCHUAN EARTHQUAKE B05301B05301

12 of 15

the Wenchuan earthquake. However, the rupture spontane-ously stops at the junction (x = −100 km) between MDS andNES in OBLIQUE before reaching the artificial faultboundary at x = −150 km. This spontaneous arrest of rupturein OBLIQUE results from a mismatch between the initialstress orientation (a small angle a = 30°) and the faultgeometry (a shallow dipping angle b = 33°) on NES. Shouldeither (but not both) of the two angles be close to 90°, therupture will not stop there. A small angle of a at thenortheast end of the Wenchuan rupture is suggested by ashear wave splitting analysis [Liu et al., 2008] and a studyon active fault slickensides [Du et al., 2009a]. A shallow dipangle at depth is also very likely, though the dip near surfacemay be steep [e.g., Burchfiel et al., 2008]. Therefore, ourmodel OBLIQUE suggests that a mismatch between the ini-tial stress orientation and the fault geometry may be the causefor rupture to stop at the northeast end in the Wenchuanearthquake.[34] This mechanism for rupture arrest can be easily un-

derstood by means of Anderson’s theory of faulting. How-ever, application of this mechanism to seismic hazardanalysis has not drawn much attention in practice. Becauseboth fault geometry and initial stress orientation can be in-ferred from geological and geophysical observations, suchas seismic imaging for fault geometry and shear wavesplitting for stress orientation, this mechanism can be used

in seismic hazard analysis to assess the rupture extent fromfuture large earthquakes.

5.2. Ground Motion and Seismic Intensity

[35] Distribution of ground motion in our modelOBLIQUE may explain some features in ground motion andseismic intensity observations from theWenchuan earthquake.Two of themmay be direct consequences of the stress rotationbetween the southwest segment and the middle segment ofthe ruptured fault in the event. First, a contour plot of PGAfrom the Wenchuan recordings exhibits a relatively largePGA zone extending from the fault trace onto the hangingwall (Z. Wang, unpublished manuscript, 2008) (http://peer.berkeley.edu/events/ 2008/2008‐18‐08_ Wenchuan‐Seminar/presentations/ZifaWang‐SFO‐public.pdf). This feature corre-lates well with the zone extending from the junction of SWSand MDS on the fault to the hanging wall in OBLIQUE(Figure 12, arrows). Second, the China Earthquake Admin-istration reported that there are two severe destruction zoneswith a lower‐level destruction zone in between in the seismicintensity distribution of the earthquake (http://www.cea.gov.cn/manage/html/8a8587881632fa5c0116674a018300cf/_content/08_09/01/1220238314350.html). This pattern cor-relates well with the pattern of two large displacement pat-ches from OBLIQUE (Figure 14).

Figure 15. Ground acceleration and velocity waveforms at four stations 1 and 10 km away from thefault on the two sides of the fault along the profile of x = −25 km from OBLIQUE (dashed lines) andTHRUST (solid lines). The stations with negative y coordinate values are on the hanging wall.

DUAN: STRESS ROTATIONS IN THE WENCHUAN EARTHQUAKE B05301B05301

13 of 15

[36] The China Earthquake Administration also reportedthat the destruction level is higher on the northwestside of the fault than on the southeast side and decayof the destruction level from the fault trace is slower onthe former than on the latter. This asymmetry can be ex-plained by asymmetric ground motion between the hang-ing wall and footwall of the westward dipping fault asshown in our models, which is a consequence of dippingfault geometry and has been documented by previousstudies of dip‐slip faulting [e.g., Oglesby et al., 2000].Contrasts in the destruction level among several cities,including Wenchuan, Beichuan, and Chengdu, can also beunderstood by peak ground motion distribution from ourmodels (Figures 12–14).

5.3. Surface Rupture and Slip at Depth

[37] Slip rake rotation near the free surface in dip‐slipfaulting has been studied by Oglesby and coworkers[Oglesby et al., 2000; Oglesby and Day, 2001]. This featureis reproduced along the entire fault length in THRUSTand along SWS in OBLIQUE in this study. An implica-tion of this result for the Wenchuan earthquake is that oneshould be cautious in inferring slip direction at depth(thus faulting style) from surface rupture, as the directionof slip vector at the surface may be significantly differentfrom that at depth. For example, right‐lateral slip of 1–2 mis observed at several locations along the southwest seg-ment of the Benchuan fault [Liu‐Zeng et al., 2009]. Thisstrike‐slip component may be mainly due to slip rake ro-tation near the free surface of pure thrust faulting at depthon the segment.[38] Another implication of OBLIQUE for the Wenchuan

earthquake in slip distribution is the difference between theextent of surface rupture and subsurface slip. Subsurface slipmay terminate earlier before surface rupture stops, resultingin a shorter subsurface rupture extent along the Beichuanfault in the Wenchuan event. We acknowledge that this canbe complicated by heterogeneous material and/or frictionalproperties, which are absent in our model OBLIQUE. De-spite this inconsistency in the extent, variations along strikein magnitude of surface rupture and slip at depth correlatevery well in OBLIQUE (Figure 10). For example, thejunction between SWS and MDS (x = 50 km) is the locationwith reduced slip at both surface and depth.

5.4. Supershear Rupture Associated with Strike‐SlipFaulting

[39] With a significant amount of strike‐slip componentsuch as that along MDS in OBLIQUE, rupture tends tobecome supershear if the fault plane is relatively smooth.Recent observations of supershear rupture mainly comefrom strike‐slip events [e.g., Bouchon and Karabulut,2008]. Furthermore, our model OBLIQUE suggests thatsupershear rupture may preferentially occur at shallow tointermediate depths (i.e., from the free surface to about10 km vertical depth), while rupture along deeper portion ofthe fault may be still subshear. Du et al. [2009b] analyzedthe temporal and spatial variation of high‐frequency(>0.1 Hz) energy radiation sources from the Wenchuan eventusing broadband seismic data of the Alaska regional net-work. They found that supershear rupture may occur duringthe second half of the rupture process. If this did happen in

the earthquake, our model OBLIQUE may provide amechanism for this phenomenon.

6. Conclusions

[40] Rotations of the initial stress field along strike canhave large effects on rupture propagation and near‐fieldground motion in a large earthquake, such as the 2008 Mw

7.9 Wenchuan earthquake. Given shallow dipping faultgeometry, rotations of the initial stress field from high an-gles between the most compressive horizontal stress and thefault trace to low angles can result in dramatic changes inrupture speed and the shape of rupture front and can evenstop the rupture. These changes in rupture propagation cancause distinct features in near‐field ground motion. Somefirst‐order features in the Wenchuan earthquake, such as thearrest of the rupture at the northeast end, may be primarilyinduced by rotations in the initial stress field. Understandingthese effects has important implications for earthquakesource physics and seismic hazard analysis in earthquake‐prone areas with complex tectonic environments.

[41] Acknowledgments. I wish to thank Y.‐G. Li and B. T. Aagaardfor helpful discussion on the topic and the Associate Editor and tworeviewers for their critical reviews of the manuscript. This research waspartially supported by USGS grant 08HQGR0048.

ReferencesAagaard, B. T., G. Anderson, and K. W. Hudnut (2004), Dynamic rupturemodeling of the transition from thrust to strike‐slip motion in the 2002Denali fault earthquake, Alaska, Bull. Seismol. Soc. Am., 94(6B),S190–S201, doi:10.1785/0120040614.

Anderson, E. M. (1951), The Dynamics of Faulting, Oliver & Boyd,Edinburgh, U. K.

Andrews, D. J. (1999), Test of two methods for faulting in finite‐differencecalculations, Bull. Seismol. Soc. Am., 89, 931–937.

Andrews, D. J. (2004), Rupture models with dynamically‐determinedbreakdown displacement, Bull. Seismol. Soc. Am., 94, 769–775.

Aochi, H., E. Fukuyama, and M. Matsu’ura (2000), Spontaneous rupturepropagation on a non‐ planar fault in 3‐D elastic medium, Pure Appl.Geophys., 157, 2003–2027, doi:10.1007/PL00001072.

Bouchon, M., and H. Karabulut (2008), The aftershock signature ofsupershear earthquakes, Science, 320, 1323–1325, doi:10.1126/science.1155030.

Burchfiel, B. C., L. H. Royden, R. D. van der Hilst, B. H. Hager, Z. Chen,R. W. King, C. Li, J. Lu, H. Yao, and E. Kirby (2008), A geological andgeophysical context for the Wenchuan earthquake of 12 May 2008,Sichuan, People’s Republic of China, GSA Today, 18(7), 4–11,doi:10.1130/GSATG18A.1.

Dalguer, L. A., and S. M. Day (2006), Comparison of fault representationmethods in finite difference simulations of dynamic rupture, Bull. Seismol.Soc. Am., 96, 1764–1778, doi:10.1785/0120060024.

Day, S. M., L. A. Dalguer, N. Lapusta, and Y. Liu (2005), Comparison offinite difference and boundary integral solutions to three‐dimensionalspontaneous rupture, J. Geophys. Res., 110, B12307, doi:10.1029/2005JB003813.

Du, H., L. Xu, and Y. Chen (2009b), Rupture process of the 2008 greatWenchuan earthquake from the analysis of the Alaska‐array data (inChinese), Chin. J. Geophys., 52(2), 372–378.

Du, Y., F. Xie, X. Zhang, and Z. Jing (2009a), The mechanics of fault slipof the Ms 8.0 Wenchuan earthquake (in Chinese), Chin. J. Geophys.,52(2), 464–473.

Duan, B. (2008a), Effects of low‐velocity fault zones on dynamic ruptureswith nonelastic off‐ fault response, Geophys. Res. Lett., 35, L04307,doi:10.1029/2008GL033171.

Duan, B. (2008b), Asymmetric off‐fault damage generated by bilateral rup-tures along a bimaterial interface, Geophys. Res. Lett., 35, L14306,doi:10.1029/2008GL034797.

Duan, B., and S. M. Day (2008), Inelastic strain distribution and seismicradiation from rupture of a fault kink, J. Geophys. Res., 113, B12311,doi:10.1029/2008JB005847.

DUAN: STRESS ROTATIONS IN THE WENCHUAN EARTHQUAKE B05301B05301

14 of 15

Duan, B., and D. D. Oglesby (2005), Multicycle dynamics of nonplanarstrike‐slip faults, J. Geophys. Res., 110, B03304, doi:10.1029/2004JB003298.

Duan, B., and D. D. Oglesby (2006), Heterogeneous fault stresses from pre-vious earthquakes and the effect on dynamics of parallel strike‐slip faults,J. Geophys. Res., 111, B05309, doi:10.1029/2005JB004138.

Duan, B., and D. D. Oglesby (2007), Nonuniform prestress from priorearthquakes and the effect on dynamics of branched fault systems,J. Geophys. Res., 112, B05308, doi:10.1029/2006JB004443.

Harris, R. A., and S. M. Day (1993), Dynamics of fault interaction: Parallelstrike‐slip faults, J. Geophys. Res., 98, 4461–4472, doi:10.1029/92JB02272.

Harris, R. A., et al. (2009), The SCEC/USGS dynamic earthquake‐rupturecode verification exercise, Seismol. Res. Lett., 80(1), 119–126,doi:10.1785/gssrl.80.1.119.

Hughes, T. J. R. (2000), The Finite Element Method: Linear Static andDynamic Finite Element Analysis, Dover, Mineola, N. Y.

Ji, C, G. Shao, Z. Lu, K Hudnut, J. Liu, G. Hayes, and Y. Zeng (2008),Rupture history of 2008 May 12 Mw 8.0 Wen‐chuan earthquake:Evidence of slip interaction, Eos Trans. AGU, 89(53), Fall Meet.Suppl., S23E‐02.

Kame, N., J. R. Rice, and R. Dmowska (2003), Effects of prestress stateand rupture velocity on dynamic fault branching, J. Geophys. Res.,108(B5), 2265, doi:10.1029/2002JB002189.

Liu, Q., J. Chen, S. Li, Y. Li, J. Wang, B. Guo, and S. Qi (2008), WenchuanMs 8.0 earthquake: The crustal velocity structure and stress field from thewestern‐Sichuan seismic array observations, Eos Trans. AGU, 89(53),Fall Meet. Suppl., U22B‐01.

Liu‐Zeng, J., et al. (2009), Co‐seismic ruptures of the 12 May 2008, Ms 8.0Wenchuan earthquake, Sichuan: East‐west crustal shortening on oblique,parallel thrusts along the eastern edge of Tibet, Earth Planet. Sci. Lett.,286, 355–370, doi:10.1016/j.epsl.2009.07.017.

Oglesby, D. D., and S. M. Day (2001), Fault geometry and the dynamics ofthe 1999 Chi‐Chi (Taiwan) earthquake, Bull. Seismol. Soc. Am., 91,1099–1111, doi:10.1785/0120000714.

Oglesby, D. D., and S. M. Day (2002), Stochastic fault stress: Implicationsfor fault dynamics and ground motion, Bull. Seismol. Soc. Am., 92,3006–3021, doi:10.1785/0120010249.

Oglesby, D. D., R. J. Archuleta, and S. B. Nielsen (2000), The three‐dimensional dynamics of dipping faults, Bull. Seismol. Soc. Am., 90,616–628, doi:10.1785/0119990113.

Ranalli, G. (1995), Rheology of the Earth, 2nd ed., Chapman and Hall,London. U. K.

Spudich, P. (1992), On the inference of absolute stress levels from seismicradiation, Tectonophysics, 211, 99–106, doi:10.1016/0040-1951(92)90053-9.

Wang, W., L. Zhao, J. Li, and Z. Yao (2008), Rupture process of theMs 8.0Wenchuan earthquake of Sichuan, China (in Chinese), Chin. J. Geophys.,51(5), 1403–1410.

Wu, X., B. Duan, and V. Taylor (2009), An OpenMP approach to modelingdynamic earthquake rupture along geometrically complex faults on CMPsystems, in Proceedings: The 38th International Conference on ParallelProcessing Workshops, edited by L. Barolli and W.‐C. Feng, pp. 146–153, IEEE Comput. Soc., Washington, D. C.

Xu, X., X. Wen, G. Yu, G. Chen, Y. Klinger, J. Hubbard, and J. Shaw(2009), Coseismic reverse‐ and oblique‐slip surface faulting generatedby the 2008 Mw 7.9 Wenchuan earthquake, China, Geology, 37(6),515–518, doi:10.1130/G25462A.1.

Zhang, Y., L. Xu, and Y. Chen (2009), Spatio‐temporal variation of thesource mechanism of the 2008 great Wenchuan earthquake (in Chinese),Chin. J. Geophys., 52(2), 379–389.

B. Duan, Center for Tectonophysics, Department of Geology andGeophysics, Texas A&M University, College Station, TX 77843, USA.([email protected])

DUAN: STRESS ROTATIONS IN THE WENCHUAN EARTHQUAKE B05301B05301

15 of 15