18
Title Dynamic Rupture of Asperities and Stress Change during a Sequence of Large Interplate Earthquakes in the Mexican Subduction Zone Author(s) Mikumo, Takeshi; Miyatake, Takashi; Santoyo, Miguel A. Citation Bulletin of the Seismological Society of America (1998), 88(2): 686-702 Issue Date 1998-06 URL http://hdl.handle.net/2433/193405 Right © 1998, by the Seismological Society of America Type Journal Article Textversion publisher Kyoto University

Dynamic Rupture of Asperities and Stress Change during a ... · Bulletin of the Seismological Society of America, Vol. 88, No. 3, pp. 686-702, June 1998 Dynamic Rupture of Asperities

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Dynamic Rupture of Asperities and Stress Change during a ... · Bulletin of the Seismological Society of America, Vol. 88, No. 3, pp. 686-702, June 1998 Dynamic Rupture of Asperities

TitleDynamic Rupture of Asperities and Stress Change during aSequence of Large Interplate Earthquakes in the MexicanSubduction Zone

Author(s) Mikumo, Takeshi; Miyatake, Takashi; Santoyo, Miguel A.

Citation Bulletin of the Seismological Society of America (1998),88(2): 686-702

Issue Date 1998-06

URL http://hdl.handle.net/2433/193405

Right © 1998, by the Seismological Society of America

Type Journal Article

Textversion publisher

Kyoto University

Page 2: Dynamic Rupture of Asperities and Stress Change during a ... · Bulletin of the Seismological Society of America, Vol. 88, No. 3, pp. 686-702, June 1998 Dynamic Rupture of Asperities

Bulletin of the Seismological Society of America, Vol. 88, No. 3, pp. 686-702, June 1998

Dynamic Rupture of Asperities and Stress Change during a Sequence of

Large Interplate Earthquakes in the Mexican Subduction Zone

by Takeshi Mikumo, Takashi Miyatake , and Miguel A. Santoyo

Abstract We investigate the spatial and temporal variations of shear stress due to the successive failures over an extensive segment of the Mexican subduction zone during a sequence of large interplate earthquakes that occurred over a period of 13 yr. For this purpose, we develop 3D dynamic rupture models incorporating a shal- lowly dipping fault located above the subducting plate. The spatial distribution of dynamic stress drop over the fault has been estimated for each of the events, through an inversion procedure using some of the previously derived kinematic fault param- eters as observational constraints.

The results revealed quite heterogeneous stress changes during these earthquakes coming from medium to high dynamic stress drop due to the rupture of a few patch- like asperities and from stress increase in between and around them. Two weak asperities located southeast of the Michoacan segment were ruptured first by the 1979 Petatlan event. The 1981 Playa Azul event ruptured two asperities in the central zone with a stress drop higher than 80 bars. The largest 1985 Michoacan earthquake resulted from the rupture of two large-size, strong asperities located at both sides of the 1981 fault zone with high stress drop of 80 to 100 bars and from another two asperities at depth. Two days after this largest event, two asperities were broken during the Zihuatanejo aftershock in the southeastern adjacent zone. Many after- shocks of these large events tend to be distributed in the zones of stress increase outside the asperities, while only small numbers of aftershocks have been observed within these asperity zones. It appears that several major asperities that existed in this extensive segment have been ruptured successively so as to fill unbroken gaps on the plate interface. Thus, the stress change left over from the previous earthquake has dominant effects on the next event in this subduction zone.

Introduction

A number of large to great earthquakes have taken place with a relatively short recurrence time along the Pacific coastal region of Mexico (Singh et al., 1981), where the Cocos plate subducts beneath the North American plate. The subduction zone extends over the entire length of about 1000 krn along the Middle America trench from the Jalisco-Co- lima regions through the Michoacan-Guerrero regions to the Oaxaca region, which appear to be segmented by the Rivera fracture zone, the East Pacific Rise, the Orozco and O'Gorman fracture zones, and the Tehuantepec ridge, re- spectively (Fig. 1) (Singh and Mortera, 1991). Most of the major earthquakes in this zone are characterized by an un- derthrusting mechanism with low dip angles and shallow source depths. The low dip of the interplate contact contin- ues to a depth of about 30 km and changes laterally some- what steeper westward beneath the Rivera plate and east- ward beneath the Caribbean plate (Pardo and Suarez, 1995). The shallow maximum depth of thrust faulting is also char-

acteristic of the Mexican subduction zone as compared with that in other subduction zones having similar age and rela- tive convergence velocity (Suarez and Sanchez, 1996).

Of particular importance from the seismotectonic point of view are the large earthquakes that occurred in the north- ern segments of the subduction zone between the Rivera and Orozco fracture zones. The segments have ruptured in a se- quence of six large earthquakes (7.0 < Ms < 8.1) over a period of 13 yr from 1973 to 1986: the 1973 Colima earth- quake (Ms = 7.5) between the Jalisco and Michoacan seg- ments; the 1979 Petatlan earthquake (Ms = 7.6) in the west- ern part of the Guerrero segment; the 1981 Playa Azul earthquake (Ms = 7.3); the 19 September 1985 Michoacan earthquake (Ms = 8.1) and its largest aftershock, the 21 September 1985 Zihuatanejo earthquake (Ms = 7.6); and the 1986 earthquake (Ms = 7.0) in the Michoacan segment (Table 1). Figure 2 reproduces a map of the rupture zones of these large earthquakes inferred from the distribution of

686

Page 3: Dynamic Rupture of Asperities and Stress Change during a ... · Bulletin of the Seismological Society of America, Vol. 88, No. 3, pp. 686-702, June 1998 Dynamic Rupture of Asperities

Dynamic Rupture of Asperities and Stress Change during a Sequence of Large Interplate Earthquakes 687

" , / , , . M E X I C. 0 6ulfofMex/co 20°N

'":'i)"":'i '".'~.' ~ $ , / ~ / 1 0 N 0 ~ T H A M E !R ~ C A 17 P L A T °%..

i WJ Ill IllllJ ,,'"" -

= , i , , , li : llllllll' , " 106°W 10.4 102 100 98 96 94°W

Figure 1. Tectonic features of the Mexican subduction zones including the Middle America trench, fracture zones, and ridges (modified from Singh and Mortera, 1991). Contours marked B indicate the depth to the upper plane of the Wadati-Benioff zone. The dotted region along tile coast shows the zone of strong interface coupling when large thrust earthquakes take place.

Table 1 Recent Major Earthquakes in and around the Michoacan Segment

of the Mexican Subducfion Zone (taken from Singh and Mortera, 1991)

Lat Lon M0 Earthquakes Date (° N) (° W) Ms (× i0 z8 dyn.cm) Mw

1. Petatlan 03/14/1979 17.46 101.46 7.6 0.55 7.62 2. PlayaAzul 10/25/1981 17.75 102.25 7.3 0.85 7.43 3. Michoacan 09/19/1985 18.14 102.71 8.1 3.90 8.05 4. Zihuatanejo 09/21/1985 17.62 101.82 7.6 0.90 7.66 5. Aftershock 04/30/1986 18.42 102.99 7.0 0.30 6.99

their aftershocks (Havskov et al., 1983; UNAM Seismology Group, 1986).

These large earthquakes have been analyzed by a num- ber of investigators from forward modeling or moment-ten- sor inversion of teleseismic, long-period body and surface waves (e.g., Singh et al., 1984; Astiz et al., 1987; Singh and Mortera, 1991; Ruff and Miller, 1994). These earlier anal- yses, mostly based on a simple point-source assumption, provided their geometrical source parameters including the source depth, fault strike and dip, and slip angle, as well as the overall source function, seismic moment, and in some cases, average slip and static stress drop. The estimated fault slxike and slip angle are generally consistent with the con- vergence direction between the Cocos-North America plates

predicted from the global plate tectonic models (McNally and Minster, 1981; DeMets et al., 1990; Ruff and Miller, 1994). Recent waveform inversions have revealed more de- tailed features of four of these large earthquakes. The spatial distribution of slip over the fault plane has been recovered from the inversion for the 1979 Petatlan, 1981 Playa Azul, 1985 Michoacan, and 1985 Zihuatanejo earthquakes, using teleseismic P waves (Mendoza and Hartzell, 1988; Men- doza, 1993, 1995), and particularly for the Michoacan earth- quake, using strong-motion velocity records with long- and intermediate-period teleseismic P waves (Mendoza and Hartzell, 1989) and near-source strong-motion records to- gether with coastal surface uplift measurements (Mendez and Anderson, 1991).

Although the previous studies have provided some basic features and kinematic fault models for each of these earth- quakes, their dynamic rupture processes and hence the change of stress state due to the successive occurrence of these events have not yet been made clear. Our main purpose in the present article is to elucidate how the spatial and tem- poral variations of shear stress could occur due to the suc- cessive failure on an extensive segment over the large-plate boundary during a sequence of large earthquakes. For this purpose, we develop 3D dynamic rupture models of the four large thrust earthquakes in this subduetion zone, incorporat- ing a shallowly dipping fault plane rupturing a part of the interface between the subducting Cocos plate and the over-

Page 4: Dynamic Rupture of Asperities and Stress Change during a ... · Bulletin of the Seismological Society of America, Vol. 88, No. 3, pp. 686-702, June 1998 Dynamic Rupture of Asperities

688 T. Mikumo, T. Miyatake, and M. A. Santoyo

(a)

19" N

1 x

\ 197~ if------.,, 4//-(7-5)

1985 (8,1)

1981 __j - 4%",... (7.3)

~'~% ~'~. . ,

i r '. I r

'1 I 1 AFTERSHOCKS 20 TO 30 SEP. 1985

M c < 3 x 20 Sep 1985

3 < _ M C < 4 ~ 3 ~ M c < 4 •

Me> '4 o Mc>__4 .

Slrong motion stations (~)

Portable stations +

1985. (7.51

I04" W ~ 3 * I02 °

&

a •

x

+ 1979

/ - . w t \

-1

,°!

(b) •2-"- o • M<~.O o M>_3.O • Portable A r r a y

BVT 6 A~celerogroph A PDE

|

Ployo AZUl L(}z(:lro Cdrden¢ls | CBZ PLV

18 o - - ,1, • 'POE UON

0L , I I I 50 km I month aftersh0ck n,oh ;~,2,72

T r e n ~

17 ° N L - - I03o loz o w

Figure 2. (a) Rupture zones of five large earthquakes in and near the Michoacan seg- ment, inferred from the distribution of after- shocks (UNAM Seismology Group, 1986). (b) Aftershock locations of the 1981 Playa Azul earthquake (Havskov et aL, 1983).

riding North America plate. In these models, some of the kinematic fault parameters previously derived from wave- form inversion (Mendoza and Hartzell, 1989; Mendoza, 1993, 1995) are used as observational constraints. From this analysis, we obtain spatially heterogeneous static and dy- namic stress drops over the fault plane of each earthquake. We will show how dynamic rupture involving that of a few asperities could develop into a large fault zone and how the shear stress increases around and between these asperities. Finally, we will examine how the stress left over from the

previous large earthquake have affected the subsequent events under the tectonic stress coming from plate move- ments.

Method of Dynamic Rupture Modeling

The method of dynamic rupture modeling has been de- scribed in Miyatake (1992), Mikumo and Miyatake (1993, 1995), Fukuyama and Mikumo (1993), and Beroza and Mik- umo (1996). To deal with dynamic rupture on a dipping

Page 5: Dynamic Rupture of Asperities and Stress Change during a ... · Bulletin of the Seismological Society of America, Vol. 88, No. 3, pp. 686-702, June 1998 Dynamic Rupture of Asperities

Dynamic Rupture of Asperities and Stress Change during a Sequence of Large Interplate Earthquakes 689

thrust fault, we basically follow the approach taken by Mik- umo and Miyatake (1993) with some improvements. Our dynamic model is specified by the spatial distribution of dy- namic stress drop and strength excess over the fault, with dynamically calculated source-time functions. We will de- rive the distribution of these dynamic parameters from the distribution of fault slip and rupture times obtained from the previous kinematic waveform inversion.

The first step in this approach is to construct a 3D dy- namic model on the dipping fault plane embedded in a hor- izontally layered half-space. The dynamic rupture propagat- ing on the fault and all displacement components can be calculated by solving the 3D wave equation with a finite- difference scheme under a fracture criterion and the appro- priate boundary conditions. This is actually a 3D shear-crack problem. The fracture criterion applied here is such that the rupture occurs at points when the increasing shear stress due to the approach of the rupture front exceeds the prescribed maximum stress level. The boundary conditions to be sat- isfied on and across the fault are (1) the initially applied and then increased, tangential shear stress drops immediately to the level of dynamic frictional stress directly above and di- rectly below the dipping fault at the time of rupture, and (2) the normal stress and normal displacement components should be continuous across the fault. Because these bound- ary conditions can be discretized into finite-difference equa- tions, six displacement components at any point directly above and directly below the fault can be solved in terms of dynamic stress drop in the case of pure thrust faulting (Mik- umo and Miyatake, 1993). Other boundary conditions im- posed are (a) the stress-free condition on the ground surface, (b) the continuity of all stress and displacement components at each of the layer interfaces (Mikumo et al., 1987), and (c) absorbing boundary conditions to attenuate reflected waves at the bottom and side boundaries of the model space (Lev- ander, 1985).

The second step is to estimate the peak shear stress just before each point on the fault ruptures. The time of rupture can be specified by the arrival time of rupture front from the kinematic waveform inversion (Miyatake, 1992). The strength excess is defined as the difference between the es- timated peak shear stress and the initially assumed stress level. However, because the peak shear stress calculated in this way is dependent on the grid spacing and the time in- crement used in numerical calculations, the estimated strength excess should be regarded as only a lower bound of its real value. In the kinematic fault models obtained for the four target earthquakes, the rupture front was assumed to propagate in a circular shape at a constant velocity. Be- cause this assumption precludes even relative estimates of the peak shear stress over the fault, we will not discuss the absolute values of strength excess in the present article.

The third step is to evaluate the distribution of dynamic stress drop from that of the kinematic fault slip. To do this, we start by estimating local static stress drop in an approx- imate way, from the formulations given by Okada (1993)

for strains and displacements at depth due to a buried point source in a half-space. By using these approximate estimates as starting values, we calculate dynamic slips over the fault, including the effects of rupture propagation and a horizon- tally layered structure, and compare them with the kinematic fault slips. The ratio between the kinematic and dynamic slips at each point is then multiplied to the previously esti- mated stress drop in the next iteration. A nonlinear iterative procedure is applied to minimize the rms difference between the kinematic and dynamic slips within a reasonably small value, and the distribution of dynamic stress drop will be obtained from a best-fitting model.

As a final check, it is hoped to see how well the dynamic models derived could reproduce the waveforms recorded in the near field, as have been done in some previous work (Fukuyama and Mikumo, 1993; Beroza and Mikumo, 1996; Ide and Takeo, 1996), because the slip velocity function and the rise time calculated from the dynamic models are some- what different from those assumed in the kinematic models. We calculated the synthetic waveforms to compare with the near-source records from the 1985 Michoacan earthquake, and the results will be given in a companion article (Mikumo et al., 1998). However, we did not calculate teleseismic syn- thetic seismograms, because some difference in the source- time function between dynamic and kinematic models will be masked in long-period, teleseismic P waveforms.

The configuration of our dynamic model is taken so as to conform the geometry of the kinematic fault model. Al- though several moment-tensor inversions gave slightly dif- ferent dip and strike for each of the earthquakes, the dip is fixed here at 14 ° for all four events, following Mendoza and Hartzell (1988, 1989) and Mendoza (1993, 1995), which corresponds to the dip of the Wadati-Benioff zone in this region; the fault strike is taken at 300 ° for three of these earthquakes, with the exception of the 1979 Petatlan earth- quake, at 293 ° , which is close to the orientation of the trench axis in this region (Mendoza, 1993). We assume the slip angle of 90 ° indicating pure dip-slip motion on a dipping thrust fault for all four events, despite that the 1985 Mi- choacan earthquake showed small right-lateral strike-slip component (Mendoza and Hartzell, 1989), because they sug- gested that the small component is not important to the over- all character of this earthquake. The fault dimension and the hypocentral depth are taken as exactly the same as those in the kinematic models.

The crust and uppermost mantle structure assumed here (Table 2) is a layered velocity model (Mendoza and Hartzell, 1989) taken from the gradient structure (Stolte et al., 1986) for the Michoacan coastal region. This is similar to the lay- ered velocity model given by Valdez et al. (1982, 1986) but somewhat different from their more refined model (Valdez and Meyer, 1996) including the subducting plate overridden by oceanic crusts. The effects of small differences in the velocity structure between the subducted slab and the over- lying mantle wedge on dynamic rupture process will be discussed in a separate article (Mikumo et al., 1998), but

Page 6: Dynamic Rupture of Asperities and Stress Change during a ... · Bulletin of the Seismological Society of America, Vol. 88, No. 3, pp. 686-702, June 1998 Dynamic Rupture of Asperities

690 T. Mikumo, T. Miyatake, and M. A. Santoyo

Table 2 Crustal Velocity Model Used in This Study

(Mendoza and Hartzell, 1989)

Vp Vs Density Top Depth Layer (krrdsec) (km/sec) (g/cm 3) (km)

1 5.80 3.35 2.68 0 2 6.40 3.69 2.78 6 3 7.00 4.04 2.85 25 4 8.00 4.62 3.00 35

they do not seem very significant if the velocity contrast is within 20%.

For numerical calculations by a finite-difference scheme, the grid spacing and the time increment should be taken so as to satisfy the stability condition for 3D wave propagation. In the present case, the fault dimension of the four large earthquakes exceeds several tens to even 180 kin, and their overall rupture duration ranges between 25 and 60 sec. To save the number of grids and the computation time, the dimensions of the model space, the crust and upper man- tle structure, and the fault size under calculation, all have been reduced proportionally to 1/10 of their actual dimen- sions. Keeping the stability condition (Vs. dt/dh <= 1/2) with the given velocity values, the time increment and the overall rupture duration can also be reduced to 1/10. Since the stress drop to give the same average slip is inversely proportional to the square root of the fault area, the dynamic stress drop from this calculation has also been reduced to 1/10 to get their correct estimates. For the 1979 Petatlan and 1985 Mi- choacan earthquakes, we take the grid spacing dh = 1.0 km and the time increment dt = 0.02 sec, while for the 1981 Playa Azul and 1985 Zihuatanejo events, dh = 0.5 km and dt = 0.01 sec. These correspond to dh = 10, or 5 kin, and dt = 0.2, or 0.1 sec, in actual cases, respectively. The di- mensions of the 3D model space in these cases are 750 × 180 × 580 km and 375 × 80 × 290 km, respectively, and the total number of grid points is about 227,000, and the total time step of 500 covers 100 and 50 sec, respectively.

Dynamic Rupture Processes of Four Large Thrust Earthquakes

The Petatlan Earthquake of 14 March 1979 (Ms = 7.6)

The Petatlan earthquake occurred in the subduction zone off the coast between Michoacan and Guerrero. This region had been quiet since 1943, but after this earthquake, a sequence of large to great thrust earthquakes successively took place in the northwestern part of this region. A number of aftershocks located by the local seismic network indicate that the Petatlan mainshock occurred on the upper interface of the subducting Cocos plate (Valdez and Meyer, 1996).

The point-source parameters of this earthquake have been estimated from P and Rayleigh waves recorded at IDA,

WWSSN, and other stations (Chad and Stewart, 1982; Singh and Mortera, 1991; Ruff and Miller, 1994). The spatial dis- tribution of coseismic slip has been inferred from teleseismic P waves recorded at the GDSN broadband network and WWSSN stations, using a linear finite-fault inversion scheme (Mendoza, 1995), under a constant rupture velocity of 3.3 km/sec and with five consecutive 1-sec boxcar source-time functions. The coseismic slip calculated over the fault di- mension of 120 × 120 km covering depths from 2 to 30 km shows a small zone with a peak slip of 70 cm near the hy- pocenter and a large-size zone of a maximum of 1.2 m far- ther southeast (Mendoza, 1995), as reproduced in Figure 3a.

Figure 3b shows the distribution of static stress drop calculated from the fault slip given in Figure 3a (Mendoza, 1995). In the corresponding dynamic model, we take the rupture starting point at a depth of 15 kin, following Men- doza (1995). The dynamic stress-drop distribution including the effects of rupture propagation is shown in Figure 3c. Although the general pattern is quite similar to that in 3b, the zone of positive stress drop is slightly more concentrated than for the static case. We notice that local stress drop of 5 to 10 bars southeast of the hypocenter generated relatively large slip with a maximum of 1.2 m and that the medium slip (70 cm) around the hypocenter comes from a small posi- tive stress drop. Zones of negative stress drop extend over the surrounding area except near the northwestern deeper section.

The estimated strength excess (not shown here) is not large but tends to have slightly higher values in the negative stress-drop zone, while it is smaller near the hypocenter and in the central portion of the high stress-drop zone. Figures 3d and 3e compare the distribution of final fault slip derived from kinematic waveform inversion (Mendoza, 1995) with that of the dynamic slip at 100 sec after the initiation of rupture calculated from the final dynamic model. The rms difference between the dynamic and kinematic slip is 5.9 cm. It seems that the dynamic model derived under a con- stant rupture velocity well reproduces the slip patterns, al- though the seismic moment 1.97 X 102v dyne cm is slightly larger than 1.5 × 1027 dyne cm in the kinematic model. It should be emphasized that relatively large slip shown here results from a dynamic stress drop of less than 10 bars. The small zones of positive stress drop may thus be regarded as real asperities.

The Playa Azul Earthquake of 25 October 1981 (ms = 7.3)

This interplate earthquake took place within the central portion of the Michoacan segment and is located between the source regions of two subsequent large events in 1985, the 19 September Michoacan and 21 September Zihuatanejo earthquakes. The centroid source parameters of this earth- quake have been estimated from body and surface waves (Dziewonski and Woodhouse, 1983; Havskov et al., 1983; LeFevre and McNally, 1985; Zhang and Kanamori, 1988; Priestley and Masters, 1986; Astiz et al., 1987; Singh and

Page 7: Dynamic Rupture of Asperities and Stress Change during a ... · Bulletin of the Seismological Society of America, Vol. 88, No. 3, pp. 686-702, June 1998 Dynamic Rupture of Asperities

Dynamic Rupture of Asperities and Stress Change during a Sequence of Large Interplate Earthquakes 691

1 9 7 9 P E T A T L A N E A R T H Q U A K E

3O

o ~ 60

r~

m 80- cl

120

( a ) SLIP DISTRIBUTION (cm)

SE N W - 2

-14

-20

26

5'0 6'0 g'o 12 0 DISTANCE ALONG STRIKE (KM)

(b) STATIC STRESS DROP ( in bars )

SE NW

O.

Izl ~ 0

I1,

:~ 80 0

<3 <

120

-I-i0-i-2-2-2-2-I-I0-i--I

-i0-2--3 2 1 1 0 0 - - I

-1 -~ 2 ~ ~ ~ o ~ o ~ o o \ 0 ~ 2 9 1 1 2 - 2 - 2 - 2 - 2 0 - i - i

1 1 3 1 3 8 - 3 0-I-I0

k0--0417-3 95 - 2 0 ~ 1

-3-i-i0-133 0-i-33 ~ i

b -i -i -i -2 --I 4 2 0-1-1 30

0 ~ 0 ~ 0 -I-I-2~1-142 0-i0

0 0 - 1 0 0 0 0 0 - 1 1 4 1 0 2

-i-I-I-100-i-i0 34 30

~o

:~ 60 0

U

120

Figure 3. Spatial distribution of dynamic and kinematic fault parameters obtained for the 1979 Petatlan earthquake: (a) slip inferred from waveform inversion (Mendoza, 1995), (b) cal- culated static stress drop, (c) calculated dy- namic stress drop, (d) kinematic slip interpo- lated from (a), and (e) dynamic slip calculated from the dynamic model.

(d) KINEMATIC FAULT SLIP (cm)

0 0 0 0 0 0 0 0 0 0 0 0 0 0

0 - 1 4 I0 12 15 17 18 16 14 12 I0 3 0

0 3 10 14 20/~'3s 48 48 44 34~s 17 i0 3

0 7 15 25 .50 77 81 71 61 49 34 6

o ,o . ..,.o ...

o ,o ~ I. 41o7) 72 4,.,., ,{o ,6 , ,

o , o / . ( , . ¢ / / . o .,,.L4V..lo._o ,,/,5 " k t L J 7 , 2 14 i, 0 , , ,o . . ,o , , , i , 3 , 3, 3, 2 , . , ,o ,o , o 2 ,

8 5 ~ 7 0 lo i , 2 , F , .0 , , , , . 4

0 . 0 ,0 I , 2 4 . . L L 2 , . . ,0 ,0 " / " i '

0 4 5 I0 14 18 16 12 8 5 12 18/ 34 23

0 2 4 5 8 I0 10 i0 7 6 15 3~0 ~ 214

0 - 0 1 1 3 5 5 4 3 6 I0 22"-24 22

0 2

0

0

0

0

0 M

0 ~

o~ 0

o

0

o

-o ~0

0 0 0 0 0 0 0 0 0 0 0 0 0 0

(C) DYNAMIC STRESS DROP ( in bars )

- 4 - 2 - 3 - 4 - 4 - 4 - 4 - 4 - 3 - 3 - 2 - I - 1

-4-2-2-3-2-1-0-0-0-0-0-0-I

-I -I --2 0

-i -I 2 6

-1 1 3 8

-1 1 3 9

--1 --I --i --I --I

-1 --i -1 -i -1

-1 -1 -1 -I -i

i i

0 30

h 4 2 3 3 1 0 0 - 1

-1 -2 -1 - i -1 -1 2 2

-0-1-2-2~2~/401~ 3

-1-i-2-2-1-043

-i-i-i-i--0 3

-I-I-1-I.-01334 t )

60 90 120

DISTANCE ALONG STRIKE (KM)

(e) DYNAMIC SLIP AT T = 1.90 SEC (cm)

0 0 0 0 O 0 0 0 0 0 0 0 0 0 0

2 3 I0~ 39 46 48 40 31~20 13 8 1

0 7 1 6 ~ ~ 4 6 ~ 36 ~25 17 iO 3

Ii

i~/3~6 I /

24 50 79 96 85 68 5~ 38 26 19 12 5

ii 5, '° "

4 5 7 9 9 9 8 0 9 24 ~ 40 26

3 . 5 ~ ~ ~ ~ ~ 71o~2~

~ o 0 6 0 0 ~ o o 6 o o ~)

0 30 60 90 120

DISTANCE ALONG S T R I K E (AM)

Page 8: Dynamic Rupture of Asperities and Stress Change during a ... · Bulletin of the Seismological Society of America, Vol. 88, No. 3, pp. 686-702, June 1998 Dynamic Rupture of Asperities

692 T. Mikumo, T. Miyatake, and M. A. Santoyo

Mortera, 1991; Ruff and Miller, 1994). Teleseismic P waves recorded at GDSN and WWSSN stations were used for source inversion to recover slip distribution on the fault. A linear waveform inversion has been applied (Mendoza, 1993), with a triangular source-time function of 1-sec duration and a constant rupture velocity of 2.6 km/sec starting from the hypocenter at a 14-km depth, for two data sets corresponding to the arrival times of small initial P phase (m2) and of major P-wave energy (MS). The distributions of fault slip over a dimension of 60 × 70 km covering depths between 6 and 23 km (Mendoza, 1993) are reproduced in Figure 4a for the M S solution. The slip indicates a concentrated distribution within a narrow zone with a radius of about 10 km located in the downdip section at depths between 13 and 20 kin. The maximum slip in the main source region exceeds 3.5 m and is about 1 m in the updip region.

Figure 4b shows the distribution of static stress drop calculated from the fault slip given in 4a (Mendoza, 1993). For the corresponding dynamic model, we included the rup- ture propagation as in the previous case. The pattern of dy- namic stress drop shown in 4c appears quite similar to the static case (4b). It is found that high stress drop exceeding 50 bars with a maximum of 94 bars is the source generating the large slip concentrated in the downdip section and that medium stress drop with 15 to 20 bars corresponds to the isolated large-slip zone in the shallower fault section. Again, these high stress-drop zones are surrounded by negative stress drop covering the remaining part of the fault. Quite large, negative stress drop can be identified between the me- dium stress drop in the updip and the high stress drop in the downdip sections. This suggests that the failure of two sepa- rated asperities caused this earthquake, and then the shear stress in this interstitial zone and the surrounding areas in- creased after the rupture. The strength excess (not shown here) has relatively large values in this zone and in the deep- est fault section just below the high stress-drop zone. These zones would have delayed rupture propagation if a nonuni- form rupture velocity had been assumed.

The distribution of dynamic slip at 35 sec after the rup- ture initiation that has been calculated from the final dy- namic model are compared with that of final fault slip de- rived from kinematic inversion (Mendoza, 1993) in Figures 4d and 4e. The rms difference between the dynamic and kinematic slips over the fault is 4.8 cm, suggesting that the present dynamic model could reproduce the slip patterns quite well. The total seismic moment calculated here is 8.7 × 1026 dyne cm, while it is 7.1 × 1026 dyne cm in the

kinematic model.

The Michoacan Earthquake of 19 September 1985 (Ms = 8.1)

This was the largest earthquake that occurred in the Mexican subduction zone since the 1932 Jalisco earthquake (Ms = 8.2) and ruptured the Michoacan segment of the Co- cos-North America plate boundary over at least 150 kin, causing heavy damage in the Valley of Mexico. The source

parameters of this unusually large earthquake have been es- timated from a number of observations mainly by body and surface waves (Anderson et al., 1986; UNAM Seismology Group, 1986; Eissler et al., 1986; Ekstrom and Dziewonski, 1986; Houston and Kanamori, 1986; Riedesel et al., 1986; Priestley and Masters, 1986; Astiz et al., 1987; Singh et al., 1988; Singh and Mortera, 1991; Ruff and Miller, 1994). Sev- eral analyses indicate that the mainshock consisted of two subevents, each with the average duration of 12 to 16 sec separated by about 25 to 28 sec, suggesting that the second subevent occurred about 70 to 95 km southeast of the first event. The average rupture velocity between the two sources ranged from 2.0 to 2.8 km/sec propagating toward the di- rection of 97 ° to 100 °. The dimension of the rupture zone has been estimated from the extent of aftershock area as about 170 × 50 km (Fig. 2a), and the average slip as of the order of 2.2 m (UNAM Seismology Group, 1986) and 4.3 m (Ruff and Miller, 1994). An attempt was also made to sim- ulate the recorded near-field displacements by a 2D cracklike model (Yomogida, 1988).

A linear least-squares waveform inversion of near- source strong-motion records and teleseismic P waves has been made (Mendoza and Hartzell, 1989) to recover the slip distribution on the fault plane with a dimension of 180 × 140 km covering a depth range of 6 to 40 km and with a hypocentral depth of 17 kin. Rupture was assumed to prop- agate at a constant velocity of 2.6 krrdsec, with a 2-sec tri- angular source-time function. The results, as reproduced in Figure 5a, indicate the existence of three major sources of predominantly dip-slip motion: (1) a peak slip of 6.5 m in a region of 80 × 55 km near the hypocenter, (2) a peak slip of 5 m in a 45 × 60 km area on the southeast portion 70 km away from the hypocenter at depths between 10 and 24 km, and (3) a peak slip of 3 m in a 30 × 60 km area at depths between 27 and 39 km (Mendoza and Hartzell, 1989). On the other hand, a frequency-domain inversion (Mendez and Anderson, 1991), using surface static displacements re- covered from strong-motion records and coastal uplift, in- dicates two areas of high slip velocity and large slip, one located downdip of the hypocenter with a slip exceeding 3 m and the other with a 4-m slip, which are separated from each other by about 100 km.

In the dynamic model that follows, we refer to the re- sults given by Mendoza and Hartzell (1989), to compare with those obtained by the same technique for the other three earthquakes. Figure 5b provides the static stress-drop distri- bution calculated from the distribution of final slip (Mendoza and Hartzell, 1989) shown in Figure 5a. Note that the dis- tribution involves three high stress-drop zones nearly cor- responding to the large-slip zones and wide negative zones separating the positive zones. The distribution of dynamic stress drop shown in Figure 5c, which included the effects of coherent rupture propagation and were calculated for a horizontally layered structure, still retains the general pattern similar to 5c. We notice that high stress drop with a maxi- mum of 100 bars near the rupture nucleation point yielded

Page 9: Dynamic Rupture of Asperities and Stress Change during a ... · Bulletin of the Seismological Society of America, Vol. 88, No. 3, pp. 686-702, June 1998 Dynamic Rupture of Asperities

Dynamic Rupture of Asperities and Stress Change during a Sequence of Large Interplate Earthquakes 693

~.~ 20

c~

Z

rn C~

1981 PLAYA AZUL EARTI-IQUAIiE

SE

(a) SLIP DISTRIBUTION (cm)

MS NW

2 8

i _ _ I

20 40

DISTANCE ALONG STRIKE (KM)

l 0

14

18

2 2

6o

h< v

5~ Figure 4. Spatial distribution of dynamic and kinematic fault parameters obtained for the 1981 Playa Azul earthquake: (a) slip inferred from waveform inversion (Mendoza, 1993), (b) calculated static stress drop, (c) calculated dynamic stress drop, (d) kinematic slip inter- polated from (a), and (e) dynamic slip calcu- lated from the dynamic model.

2O

0 C~

4 0 (D

5-, co

60

SE (b)STATIC STRESS DROP ( in bars ) NW

o 2 - 1 - 2 3 3 1 - 4 - 6 - 6 - 4 - 2 0 - 3 - 3

- 1 - 3 ~ ~ - 4 ~ 3

- 1 - 4 - 3 - 6 - 4 3 I-6-4

- 4 - 3 0 - -10-9-7

-3-30-4-21 -6-8-6

--3-217--7 - 10 -173

-3--2--411-3--1-2-15-13-9--4-3

- 1 - 3 - 4 - 3 0 " 0 - 5 - 5 - 3 - 3 - 3 - 2 - 2

- 1 - 1 - 3 - 3 - 1 0 2 - 4 - 3 - 3 - 2 " - 1 - 1

v

2Z

(c) DYNAMIC STRESS DROP (.in bars )

-s -4(o o o -\-~ -9 -~ -6 -5 -~ -5 -5

-5 -4 k~o 1 o "Go -Tq ~-o -5 -G -s

--5 -5 0 0 ii 19 15 8 0 ~5 --5

--5 --5 --5 --5 0 15 14 15 14 18 14 --5 ~5

-5 -4 - --6

-5 -3 -32 -33 -20- 3 / - ,

--4 --3 0 - 23 45 " 75 53 --2 --5 -6

-5 - 3 - 4 ~ ~ - ~ ' ~ ok,t_ o/~-~ ~ ~ - 1 , - 1 4 - ~ 3 - 1 1 - ~

-5 -4 -4 -6 -8 --5 -14 -14 -14 -ii -13 -II -S

-5 --4 -6 -6 -5 --5 --13 -i0 -10 -i0 -7 --7 - 5

--5 --4 -6 -6 -5 --5 --i0 -10 -10 -i0 -7 --7 --8 £ J

0 20 ,40 60

DISTANCE ALONG STRIKE (KM)

0

00"

o

o

o

M ~ 2 0 o

~-~ o c~

o Z;

o o c~

40 o- ~J U

o <

o 09 H r~ 0

60

(d) KINEMATIC FAULT SLIP (em)

0 0 0 0 0 0 0 0 0 0 0 0

0 3 15 20 20 17 15 20 22 20 19 5 0 " 0 6

28 ~ ^ ~ ~ ~ o0 ~ 32 ~3 44 22 i0 3 0

0 O 8 15 ~ ~ 3 2 47 17 5 0

0 0 i0 17 51 82 94 10 29 10 0

0 5 19 32 5 15 0

0 I0 18 50 42 20 18 I~ 37 34169 ~4 18 0 [ /

35 18 50 39 47 42 16 0

2 12 26 7 0

3 13 29 7 0

,. 7o, 27 o //

17 ,o o 211 24

0 7 15 33 4 ~ ~ 0 18 8 6 3 0

0 - 0 2 12 16 21 38 25 22 18 12 2 0 0 G

0 0 0 2 7 17 20 12 i0 9 6 I 0 0 0

0 0 0 0 2 7 10 6 3 2 0 0 0 O - 0 2~

0 0 0 0 0 0 0 0 0 0 0 0 0

(e) DYNAMIC SLIP AT T = 35 SEC (cm)

0 0 0 O 0 0 0 0 0 0 0 0

0 4 13 17 22 21 22 25 26 23 13 7

0 5 15 24 31 4~4925 13 / /~. \

I 5 io ~ 5 84 I~2 9 ~ 18 39 55 84 102 98 80 49 23

1 6 12 %~ 21 48 79 90 97 I~I~_~03 82\~3 /

3 I0 17 32 6~ ~ / ~ ~ i 8 5

3112143,227 0o)°° so

5 l i 24 33 22 32 1 4 4 " k 9 0 ~ 5 0 " 55 41

5 12 26 33 29 58 27

5 14 30 42 63 25

4 i3 30 50 74 5 22

i 7 1, 30 03 s) -29 13 5

0 4 9 17 24 38 32 30 21 12 4 i

0 2 5 9 14 20 12 12 ii 6 3 0

1 2 2 5 8 i0 6 6 4 0 0 0

o o o ~ o o o ; o o o

20 40

DISTANCE ALONG STRIKE (KM)

0

4 o

6 o

11 o

15 0

16 o

o

o

o

0

o

o

2 o

o o

o o

o o

i o

60

Page 10: Dynamic Rupture of Asperities and Stress Change during a ... · Bulletin of the Seismological Society of America, Vol. 88, No. 3, pp. 686-702, June 1998 Dynamic Rupture of Asperities

694 T. Mikumo, T. Miyatake, and M. A. Santoyo

5:f

el.. 30

Z

6o o E3 LLI © Z 90 .< F- 09 23 120

1985 MICltOACAN EARTHQUAKE

(a) BLIP DISTRIBUTION (cm)

SE NV

30 60 90 120 150

DISTANCE ALONG STRIKE (KM)

M v

b ~

C~

40

Figure 5. Spatial distribution of dynamic and kinematic fault parameters obtained for the 1985 Michoacan earthquake: (a) slip inferred from waveform inversion (Mendoza and Hart- zell, 1989)--solid circles marked by PLA and ZIH indicate the hypocenters of the 1981 Playa Azul and 1985 Zihuatanejo earthquakes, (b) calculated static stress drop, (c) calculated dy- namic stress drop, (d) kinematic slip interpo- lated from (a), and (e) dynamic slip calculated from the dynamic model.

~E

(b) STATIC STRESS DROP ( in bars )

o ...... , ..... 6 .... , . . . .

| 2t...5'

--720212014 6 4 618 25 122412 -4 ~3

-3 22, 2625 2 610 27 ZO 302 16 1 1

60 2 -51O 27 51919 8

/ 3 --5 -6 -~ ~15 -14 ~12 --12 --6 ~5 -9 --9 -9 --10 --5 -5 -6 -2

90 2 -i -5 -7-12 -9 -4 -12 -9 -5 -3 2 ° -4 -5 -i -I 0

~ - 5 ~ - 5 -~ -4 -i0 -G ~3 -2 -I -i -I - 1 o 1

120 J--3 20 28 15 5 1 0 1 1 1 1 1 2

[ --2 212112 6 1 1 2 2 3 2 3 6

(C) DYNAMIC STRESS DROP ( in bars )

17- ...................................... ~ -~ ~ q - ~ ~ . . . . ~-~ -6 -~ -S -s -~ -~ o i0 0 -8 0 - -I -5 -2 o

- g ~ - 5 -3 0

-S -5 -2 0

--i ° :'6 2

-i -7 -i0 -ii -19 13 -9 -13 -6 -~ ~iI -i0 -12 -12 -5 -3 -7 -3 0

0 -3 -8 -I0 9 -a -5 -I0 -~ -5 -2 0 -I -4 -5 -~ 0 0 0

-G -5 - 1 5 -7 -5 -3 -2 -2 -i -1 0 0 0 2

L- 4

60 120 180

DISTANCE ALONG STRIKE (KM)

NW

6

40

(d) KINEMATIC FAULT BLIP (cm)

0 o o o o 0 0 0 o o o o o 0 o o o

0

~o v

120

5047353942333430433941403"7454835272422

43 63 67137 i~8 i~ IF 129117 4 72 72 95 97 ?9 49 28 22 Zl

2431 {i ~1891~(9101~977216 141711514015145~185181 78'2 ~ ~ 9 5 "92 373017191919

52 O1169 IUl 714235

94 68 i~2~958/~713327269582~ %06130 . . . . 145 ii0 i19 ~ 1 2 77171823

5949751Q0~94" 7455255~ 6459645° 515842{ 172124

4"/ 546275749090423950687667484043322929

194780 9~6144~13~7~ 474650555449434]" 373533

178279 ~ 5 ~ ~ 7 8 7] i0166554951494743403837 \ 26/{33~oo s~\93 G~ so 5~ 5~ 48 ~s 42 43 4o

96575592238 Q~/846051 ~94347484345 ~345 o:2

0 0 ~ 0 o o o o o o o o o o o o o o o 0

(e) DYNAMIC SLIF AT T = I00 SEC (cm)

o o o o o o o o o o o 0 o o o o

323931 43 49353534394433293239 433225

3 6 4 ~ 75759390744730

465566 ~ 8 2 g161 37374451 g76143384629

195277106137 183104444245575451464440

22758490 I1520621168107476156 $449464239

2013914082471031821801268256434846434039

86673552146 I06~65042233939393533

o 0 0 0 0 0 0 0 o o o ~ o o o 0

80 ]20

DISTANCE ALONG STRIKE (KM)

0

2329 O

2318 o

18 0

16 0

29 0

31 0

30 0

19 0

1719 0

2427 o

3632 o

37 35 o

37 3~ 0

4047 o

47 45 o

o o

180

o

"0 8

o

o

0

o

0

0g 0

0 ~ 0 N o

o

o

o

'o40

Page 11: Dynamic Rupture of Asperities and Stress Change during a ... · Bulletin of the Seismological Society of America, Vol. 88, No. 3, pp. 686-702, June 1998 Dynamic Rupture of Asperities

Dynamic Rupture of Asperities and Stress Change during a Sequence of Large Interplate Earthquakes 695

an extremely large slip exceeding 6 m at depths between 10 and 18 kin. This may be the failure of the largest asperity at the first stage. The second subevent with a peak slip of 5 m located about 70 km southeast of the first one came from local dynamic stress drop larger than 80 bars, and the third subevent occurred with a stress drop of about 40 bars in the downdip fault section, giving a peak slip of 3 m. While the first and second high stress-drop zones are separated by a zone of small positive stress, the total area extending over 140 km is surrounded by negative stress-drop zones. The third medium stress-drop zone in the deepest fault section is isolated by negative stress drops from the shallower high stress-drop zones. The negative stress drop may be a natural consequence that could occur outside the positive, high stress-drop zone and the peripheral zone of the fault area.

The estimated strength excess (not shown here) has rela- tively large values in the shallow section above the first high stress-drop zone, the mid-depth southeast of the second one, and the deep zone below the third one. On the other hand, it has lower values between these zones, including around the hypocenter. This would suggest that the rupture starting from the hypocenter propagated southeastward to the second high stress-drop zone without receiving strong resistance. Figures 5d and 5e show the distribution of the kinematic slip interpolated from the results by Mendoza and Hartzell (1989) and the dynamic slip at 100 sec after the initiation of rupture calculated from the present dynamic model. The rms difference between the kinematic and dynamic slips is 7.7 cm, indicating that the slip patterns in the two models are essentially similar. It should be mentioned that large slip exceeding 160 cm can also be observed between the first and second sources where only small stress drop occurred, and that medium slip between 30 and 60 cm occurred over the negative stress-drop zones.

The Zihuatanejo Earthquake of 21 September 1985 (Ms = 7.6)

This earthquake occurred 2 days later as the largest af- tershock immediately to the southeast of the source region of the 19 September earthquake (Fig. 2a). The centroid source parameters determined by several investigators (Dziewonski and Woodhouse, 1986; Eissler et al., 1986; Priestley and Masters, 1986; Riedesel et al., 1986; Ruff and Miller, 1994) were almost the same as those of the main- shock. The dimension of the ruptured area and the average slip have been estimated as about 60 km x 30 km and 3.3 m (UNAM Seismology Group, 1986). Both teleseismic P and SH waves well recorded at GDSN stations were used in the waveform source inversion (Mendoza, 1993). The inversion was made with the same source-time function and the rup- ture velocity as in the case of the 1981 Playa Azul earth- quake. Figure 6a reproduces the distribution of fault slip (Mendoza, 1993) over a dimension of 90 X 90 km with the hypocentral depth of 20 km. The slip occurs in a circular zone with a radius of about 30 km extending over depths between 12 and 26 krn. The maximum slip exceeds 1.7 and

2.0 m for the updip and downdip portions southeast of the hypocenter.

For the dynamic model, we take a slightly smaller fault size of 70 X 80 km than that of the kinematic fault model, because fault slip in the southeast and downdip sections has very small values. This would not give significant effects in estimating stress-drop values. Figure 6b gives the distribu- tion of static stress drop calculated from that of final fault slip (Mendoza, 1993) shown in 6a. The corresponding dy- namic stress-drop distribution, which includes the effects of rupture propagation, is shown in Figure 6c. The pattern of dynamic stress drop shows somewhat different features par- ticularly near the upper and lower edges of the fault zone. This may be due to an edge effect during dynamic rupture propagation. It is found that medium stress drop higher than 20 bars in two patches yielded fault slip larger than 170 cm in the central zone of the fault, and hence, these asperities may be the source of this earthquake. This nearly circular, positive stress-drop zone is again surrounded by negative stress drops covering the remaining part of the fault.

The estimated strength excess (not shown here) has rela- tively high values in part of the updip section just above the positive stress-drop zone and in the downdip section cov- eting part of the central zone. Figures 6d and 6e compare the distribution of final slip obtained from the kinematic in- version (Mendoza, 1993) with that of dynamic slip calcu- lated from the present dynamic model. Although the abso- lute values at several points slightly differ from each other, their general slip patterns have similar features. The rrns dif- ference between the dynamic and kinematic slips is 5.8 cm. The seismic moment 1.64 X 1027 dyne cm in the dynamic model is slightly larger than 1.35 X 1027 dyne cm in the kinematic model.

Uncertainties

The dynamic models for the four earthquakes consid- ered here have been derived with some observational con- straints consistent with the corresponding kinematic fault models (Mendoza and Hartzell, 1989; Mendoza, 1993, 1995). The nonlinear iterative inversion procedure used here implies that some uncertainties involved in the calculated kinematic slip would affect almost linearly the estimates of dynamic stress drop. The kinematic waveform inversion mentioned here assumes a constant rupture velocity. Be- cause the rupture time has a strong nonlinear effect on the recorded waveforms, it is possible that this simplified as- sumption would have affected the kinematic fault slip to some extent and, hence, the dynamic stress drop. Also, a simple source-time function with a constant rise time was used in the kinematic inversion, while the slip-rate source function in dynamic models has nearly an inverse square- root time dependence depending on the location on the fault. Actually, waveform inversion incorporating the dynamically derived slip-rate time functions have yielded a more con- centrated distribution of dynamic stress drop than that de- rived from conventional kinematic inversion (Beroza and

Page 12: Dynamic Rupture of Asperities and Stress Change during a ... · Bulletin of the Seismological Society of America, Vol. 88, No. 3, pp. 686-702, June 1998 Dynamic Rupture of Asperities

696 T, Mikumo, T. Miyatake, and M. A. Santoyo

A

zo

2: N: 40 O

r.) Z 6O . <

O3

8 0

SE

1985 ZIHUATA/NEJO EARTHQUAIf~E

(a) SLIP DISTRIBUTION (cm)

NW

t -

O ZO

i , I . . . . I •

ZO 60 DISTANCE ALONG STRIKE (KM)

12

1 6

C I

8 ~

9O

Figure 6 . Spatial distribution of dynamic and kinematic fault parameters obtained for the 1985 Zihuatanejo earthquake: (a) slip inferred from waveform inversion (Mendoza, 1993), (b) calculated static stress drop, (c) calculated dynamic stress drop, (d) kinematic fault slip interpolated from (a), and (e) dynamic slip cal- culated from the dynamic model.

M ~ 2 0

O

~40 o

6 0

( b ) STATIC STRESS DROP ( in bars )

S E NW

9 0 0 0 ° 1 4 5 5 8 ° 5 3 0 0

0 16 ZO 23 20 13 3 3 4 - ~

-I -3 -6 9 8 3 2~ 18 6 ] 6 ii 8 3 -

..... 6 / 3 Ii<:~9 186 : I: 16 1 2 8 7 -

-2 -2 -31 ~ 1o 15 5 5 13 I~ 16 19 3 5

-3 0 - i/ 10 9 7 ~ 9 12 13 14 25 4 7

-3 - i 9 9 9 3 18 13 12 15 1B 18 3 0 ..-3

-2 -4 -41 8 ~ / ~ 8 3 18 11 13 11 13 -2 ~5 -4

0 -i -I o -8 -9 17 13 14 11 3 -5 -6 -3 -I

0 o -1 ! -2 -5 -6 2 5 8 6 I -I -5 -i o

~0 9 9 0 1 3 -4 -6 -~ l 3 5 2 -i 0

o,o o

(c) DYNAMIC STRESS DROP ( in bars )

-I -i -i -i -i -7 -B -lO -Ik -I] -iG -5 -6 -2 -2 -2 -i

~0 0 -3 -14 r7 -14 7 7 0 o 0 -8 0 0

-1 ¢ -8 o 7 5 1 4 ° 5 0 -9 -5 -1

i --2 - 3 - I 0 2 II ~ 9 S 1 8 "~ 0 - 5 - 3

[ - 4 0 - 2 1 s 9 7 7 9 s 84 14 18 Z 9 { 4 6 o / -5 -3 r

-4 - 2 0 7 9 3 20 14 13 18 22 26 3 0 -6 -6 -2

-2 -7 -~ g 8 23 21 13 18 13 13 -3 -8 -8 -2 -I

o -] -4 -3 -14 -19 17 17 19 Ii I -14 -13 -6 -2 0

0 ° 0 - 2 3 - 9 - 1 2 - 1 8 ~17 - t 0 - 1 - 2 - 2 0 0 0

o o 9 o -I -2 -3 -5 -8 -4 -i -2 -4 ~i o o o

0 20 40 60 80

DISTANCE ALONG STRIKE (KM)

- 1 2

28

0 0-

9

o

o

~ 2 0 0 -

o

t o cl

°

0 4 0 0 -

o

o

8

(d) KINEMATIC FAULT SLIP (cm)

o o o o o o o o o o o 9 o o o o o

'-o, 12 o

o

o

o

o

o

° ~

8

0

0

0 9 9 9 9 1 16 15 12 16 15 19 9 0 0 9 9

8/ 2o 68o° 9 O 0 0 O 2 53

o o o . . . . .

1 27 71 124 157 170 152 128 iII 96 69 38 5 O 0

0 9 9 4"7 184 i0 0 O

44 96 118 133 150 168 178 180 139 177 i09 75 38 7 0

8 8 ° ,.., ,1 o66 , o o o o /

o o ° o o i i ~ 28 1 o o o o

0 0 0 0 9 0 9 3 5 1 6 3 1 8 2 2 7 4 0 0 9 9

0 9 o 9 9 9 8 9 8 1 9 1 8 6 9 9 o 8 8

8 9 9 9 9 0 0 9 0 9 9 0 9 8 9 0 0 8

o

o

-o 28

D O O 0 O 0 O O 0 0 0 0 0 0 9 0 0 O

(e) DYNAMIC SLiP AT T = 50 SEC (em)

o o o o o o o o o o o o o o o o o o o

O 1 3 5 7 8 8 15 17 16 10 7 14 7 8 4 1 0 0

7 8 7 0 42 GG 79 90 @I 76 51 22 14 5 3 0 O G

0 8 6 1 22 55 107 131 143 128 i~i 73 54 36 8 4 0

0 5 O 44 78 22 151 167 151 132 116 99 75 47 2 2 4 0

2 \ \ 0 2 6/ / (I652<II 12 ~ / 0 ~ 6 2 12 ~ 6 2 105 158 172 133 151 140 133 118 92 59 7 i 0

0 8 i0 76 123 162 159 158 161 !64 162 140 i06 36 40 3 0

8 9 97 85 106 120 168 172 169 1~3 168 48 £B 54 1O 0 8

o 2 o 198 0 3 o

0 5 8 3 0 4 55~71 75 ~54 36 0 1 3 4 0

0 4 4 8 2 0 4 2 6 42 38 29 4 1 2 3 9

0 1 2 2 3 2 2 8 0 23 25 5 0 1 1 2 0

o o o o o o o o o 6 o o o o o o o 9

20 40 60 80

DISTANCE ALONG STRIKE (KM)

Page 13: Dynamic Rupture of Asperities and Stress Change during a ... · Bulletin of the Seismological Society of America, Vol. 88, No. 3, pp. 686-702, June 1998 Dynamic Rupture of Asperities

Dynamic Rupture of Asperities and Stress Change during a Sequence of Large Interplate Earthquakes 697

Mikumo, 1996). Some difference in the velocity structure between the subducting plate and the overriding mantle wedge could also slightly affect dynamic stress drop on a dipping thrust fault. Thus, the absolute values and even the distribution pattern of dynamic stress drop obtained here could be subjected to minor changes (say, about 10%) if the above conditions are taken into consideration. However, it would be interesting to compare the general pattern of stress drop with aftershock distribution from each of the four earth- quakes analyzed here, and also to discuss the spatial and temporal variations of shear stress over an extensive segment of the plate boundary.

drops (Figs. 5b and 5c), these aftershocks may have been caused by stress increase or the deficit of stress decrease. In the zone just southeast of the epicenter and in the rupture zone of the 1981 Playa Azul earthquake, there was large stress drop, which could not generate many aftershocks. Af- tershocks of the second event occurred southwest of the mainshock and mainly in the southeastern region, while al- most none was observed around the epicenter. Again, it ap- pears that the increase of stress after the mainshock in the southwest and southeast regions (Figs. 6b and 6c) have gen- erated many aftershocks, while the medium stress drop around the epicenter did not generate aftershocks.

Stress Drop and Aftershock Distribution

Now, we compare the general pattern of stress drop dur- ing each earthquake with the distribution of its aftershocks. Because the earthquake epicenters in this region determined only from teleseismic observations such as given in the ISC and PDE catalogs are often displaced by 30 to 40 km north- east of those from local and regional networks (e.g., Hav- skov et al., 1983; Ruff and Miller, 1994), we refer only to the epicentral locations determined by the local seismic net- work.

The locations of aftershocks of the 1979 Petatlan earth- quake have been determined (Valdez and Meyer, 1996; not shown here) by the mobile seismic stations. It does not ap- pear that these aftershock distributions can be correlated with any stress-drop zones (Figs. 3a and 3b), unlike the case of the following three large events. This might be due to much smaller stress drops (<10 bars) in this case, Figure 2b shows the locations of the mainshock and aftershocks of the 1981 Playa Azul earthquake (Havskov et aL, 1983), which have been obtained for the first 6 days by a portable seis- mograph array and a strong-motion station in this region. The aftershocks cluster in the eastern and western sides of the zone delineated by a dotted curve, while there is a gap just north of the mainshock epicenter. The gap corresponds exactly to the high stress-drop zone in the downdip section (Figs. 4b and 4c) just below the hypocenter, suggesting that the large drop of stress could not generate aftershocks in this zone. On the other hand, it is noticed that the increase of stress in the eastern and western sides of the foregoing zone was capable of generating these aftershocks.

Figure 2a gives the locations of the mainshock and af- tershocks of the 1985 Michoacan and Zihuatanejo earth- quakes (UNAM Seismology Group, 1986). All the locations have been obtained from a portable seismograph array and strong-motion stations in this region for the data during 10 days after the second event. Aftershocks of the first event are distributed nearly in the northwestern section of the mainshock epicenter as well as in the southeastern end re- gion, but only small numbers of aftershocks have been lo- cated in the central zone. Because the region northwest of the epicenter and near the southeastern end zone had en- countered negative stress drop or only small positive stress

Failure of Asperities and Stress Change over an Extensive Segment

A sequence of the four earthquakes in the Michoaean- western Guerrero segment appear to have taken place on parts of nearly the same interface between the subducting Cocos plate and the overriding North American plate. Fig- ures 7 and 8 accommodate the fault planes of these earth- quakes projected onto this interface. Because the initial shear stress prior to the 1979 first event may not be uniformly distributed on the interface, we are looking only at the change of stress state due to the sequence of these earth- quakes. Figure 7 shows the zones of positive stress drops higher than 10 bars. Letters PET, PLA, MIC, and ZIH indicate the rupture starting points (i.e., the hypocenters) of the 1979 Petatlan, 1981 Playa Azul, 1985 Michoacan, and 1985 Zih- uatanejo earthquakes, and contours with different symbols show dynamic stress drop during these events. We see that these stress-drop zones are distributed like patches over an extensive area but overlapped only on a small area of the 1981 Playa Azul and 1985 Zihuatanejo fault zones. If, how- ever, we take stress-drop zones with higher than 15 bars, these zones are not overlapped but just adjacently located. On the other hand, fault slips from the four earthquakes over- lap in several areas: the northwestern part of the fault zone of the 1979 Petatlan earthquake reslipped during the 1985 Michoacan and Zihuatanejo earthquakes, the fault zone of the 1981 Playa Azul earthquake also slipped again during the two 1985 events, and the southeastern part of the fault zone of the Michoacan earthquake appears to have reslipped during its largest aftershock (see, Mendoza, 1993, 1995).

If we define here an asperity as part of the fault zone having medium to high positive stress drops, for example, higher than 10 to 15 bars, we identify one asperity for the Petatlan earthquake, two asperities with one having 80 bars stress drop for the Playa Azul earthquake, four large asper- ities with two having 80 to 100 bars stress drop for the Mi- choacan earthquake, and two asperities for the Zihuatanejo earthquake, over this segment of the plate boundary. Of par- ticular interest is the central fault zone of the Playa Azul earthquake, where the stress dropped by more than 80 bars. During the 1985 Michoacan event, two large-size asperities located at both sides of this zone ruptured, and the central

Page 14: Dynamic Rupture of Asperities and Stress Change during a ... · Bulletin of the Seismological Society of America, Vol. 88, No. 3, pp. 686-702, June 1998 Dynamic Rupture of Asperities

698 T. Mikumo, T. Miyatake, and M. A. Santoyo

° °

. ~ 9 q c3 ?

tO .•

: : PET

$60 ° E

I I 1 9 8 1 P - - I - - 1 9 8 5 M-

I ..... 1985 Z.-- . . . . ~ . . . . -i .--;o.

/ " / " ~ - 1 " f , . j . . I

/ -"" "" I o [ ""--- "PLA t" / " - e ",Q~

I ~"-" 3 0 '

• / ' Z I H ' - - ; ' . " ' ' ' , ' / ~ . / ~- ~0-" i ("3,° . , ~ . J \ \

o

i~- ° "

. . . . . . . . . . . . ..'~ _ .J

"t0

• .~

• • " _lO

STRESS DROP ( > i0 bars

DURING 1979 - 1985

1986 A

< ........ 180 km ......... >

Figure 7. Zones of positive stress drop higher than 10 bars projected on the fault planes of four large, thrust earthquakes: 1979 Petatlan, 1981 Playa Azul, 1985 Mi- choacan, and 1985 Zihuatanejo earthquakes. Contours with different symbols represent dynamic stress drop (in bars) during the four earthquakes. Small solid circles indicated by PET, PLA, MIC, and ZIH are the hypocenters of these earthquakes.

N60"W

i

1

zone did not see any more stress drop but still slipped due to the rupture of the two asperities. Thus, reslip occurred on several zones, and new slip took place even in negative stress-drop zones during these earthquakes. These fault slips may be regarded as free slips forced by the failure of a few major asperities with medium to high stress drop. The fore- going definition of "asperity" differs from that often used in previous studies depending on fault slip or moment re- lease (e.g., Kanamori, 1981; Ruff and Miller, 1994)• Because these asperities appear to be located near the northern ex- tension of the Orozco Fracture Zone, it is likely that all four large earthquakes might be related with the subduction of this fracture zone with the Cocos-North American plate con- vergence, as have been suggested by Ruff and Miller (1994)•

Figure 8 shows the contours of zero stress drop sepa- rating the positive and negative stress-drop zones or the zones of stress decrease and increase after each of the four events (see notations in figure caption). Although these zones overlap each other and appear rather complicated, we notice some features. The increase of stress due to the 1979 Petatlan earthquake ( + pe) had been prepared in part of the 1985 Zihuatanejo fault zone, the stress increase due to the 1981 Playa Azul earthquake ( + pl) should have been iden- tified around the western large-size asperity of the 1985 Mi-

choacan event, and the increase of stress due to the Michoa- can event ( + m) had appeared in the eastern part of the Zihuatanejo fault zone. One of the large aftershocks, the 1986 event (Ms = 7.0), took place in the northwest zone ( + m) outside the major asperity of the Michoacan earth- quake. Thus, there is the possibility that the increase of stress due to these large events could contribute, to some extent, to trigger the subsequent earthquakes.

Temporal variations of shear stress at five selected points B, D, E, PLA, and ZIH during the 8 yr from 1979 to 1986 are shown in Figure 9. During the interseismic periods, the shear stress should slowly increase due to the loading of tectonic stress coming from the convergence between the Cocos and North America plates• If we take into consider- ation the rate of convergence ranging between 5.5 cm/yr (near 104.5 ° W) and 8.0 cm/yr (near 91.5 ° W) (Singh and Mortera, 1991), assuming the width of strong interface cou- pling to be about 60 km (see Fig. 1), the secular rate of stress increase would be about 0.35 to 0.50 bars/yr, which is im- plicitly included in broken lines in Figure 9. Small rapid rise of the stress just before its large drop at the time of the earthquakes (2), (3), and (4) indicates the strength excess due to the approach of the rupture front toward these points. It is interesting to note that the stress level had increased to

Page 15: Dynamic Rupture of Asperities and Stress Change during a ... · Bulletin of the Seismological Society of America, Vol. 88, No. 3, pp. 686-702, June 1998 Dynamic Rupture of Asperities

Dynamic Rupture of Asperities and Stress Change during a Sequence of Large Interplate Earthquakes 699

°

+ pe

+ p e

. 1 9 " / 9 ? - -

. - • ' ' " - - ' " . . . . . . "Sg0 ° E " . , ' " = , ' , - 1 9 8 1 P k j • 1 9 8 5 M

• • ' • : + IIl I f ~ ~----~ I

• 7 - - - 1 9 8 5 Z . . . . . . . . . . . t ~ 7 ~ x I + m • .0"" ' + z •" " . ; . ' - ' " " -0 . / I ~ - J - pl ~ + m

." • e ~ , h • , . / I + n ] / f \ - x • . " t' " . 0 - z . \ p l / .

- p e ' . • / - pe . t O " e ' - " P L A ~, - m

" 1 " ] - z ; / " Z I t t - m 1', - p l / M I 5

• % 0 ~'\' 0 - " ~ B p l

"... + z ; + m ~

"0 + m • p e - p e ',

' • • 0 " . - " " ' '

. . . . . . . . . . . . . - " < 0 / -

- rI1 STRESS CHANGE

DURING 1979 - 1985

N80*W

o

/' I

$

< ........ 180 km ....... ->

Figure 8. Contours of zero stress drop separating the positive stress drop (stress decrease) and negative stress drop (stress increase) zones• + pe and -pc: zones of stress increase and decrease due to the Petatlan earthquake; +pl and -pl : zones of stress increase and decrease due to the Playa Azul earthquake; + m and - m : zones of stress increase and decrease due to the Michoacan earthquake; and + z and - z: zones of stress increase and decrease due to the Zihuatanejo earthquake. For symbols PET, PLA, MIC, and Z/H, refer to Figure 7.

an appreciable extent at three points B, E, and ZIH several years prior to the largest 1985 Michoacan earthquake.

Discussion

We would like to point out that several large-size as- perities have been ruptured successively in time so as to fill unbroken gaps on the plate interface in and near the Mi- choacan segment• There remain a few questions, however, as to why the 1981 Playa Azul fault zone ruptured as the second large event following the 1979 Petatlan earthquake, while the 1985 Zihuatanejo fault zone just adjacent to the first event ruptured at the fourth stage, and why the dynamic rupture from each of the large earthquakes has been arrested somewhere between the adjacent fault zones. One possible explanation would be that there was long-wavelength vari- ations of the coupling strength of the interface between the subducting and overriding plates, which might be attributed to some geometrical configurations, with the initial shear stress being distributed just below the strength level. Sup- pose that the average strength would be the lowest in the

Petatlan region, the second lowest in the Playa Azul region, the second highest in the two Michoacan asperity zones, and the highest in the Zihuatanejo region• It seems likely that the gradual increase of tectonic shear stress due to the plate sub- duction plus the rapid stress increase due to the dynamic failure of major asperities reached the average strength in the four regions one after another as time progresses• It is possible, however, that if the stress increase solely from the asperity failure cannot overcome the higher strength near the boundaries of the adjacent fault zones, then the dynamic rupture would be arrested there• In this case, the high strength there would work as barriers (e.g., Aki, 1979)• The estimates of the strength excess mentioned in this article are not sufficient to solve the above-mentioned problems• These explanations are only speculative at the moment• More theo- retical considerations and numerical experiments incorpo- rating long-wavelength lateral variations of strength and stress with their short-wavelength fluctuations (e.g., Mikumo and Miyatake, 1983) might help in understanding these problems as well as temporal and spatial distributions of aftershock generation•

Page 16: Dynamic Rupture of Asperities and Stress Change during a ... · Bulletin of the Seismological Society of America, Vol. 88, No. 3, pp. 686-702, June 1998 Dynamic Rupture of Asperities

700 T. Mikumo, T. Miyatake, and M. A. Santoyo

STRESS CHANGE

- - ? - - - B

--? -- -- pLA - - I

5 0 b a r s

v

? - _

?

- L _ _ ?

- - ? - - - - D

- - ? - - - - E . [

I ?

/

? . . . . . [ I ~ ?

NO I - - ? I xI0..28 (3) dynB,cm

2 (4)

1 (t) t2) (s)

L ~ i l ~o ] ~-~ L ~o~ I ~ ~ ! ~_ ~.l-

Figure 9. Temporal variations of shear stress at five selected points, B, D, E, PLA, and ZII-I on the fault zones (see Fig. 8) during 1979 to 1986. Earthquake numbers (1), (2), (3), (4), and (5) refer to Table 1. M0 indicates the seis- mic moment of each earthquake.

Conclusions

We have investigated the spatial and temporal variations of shear stress due to the successive failures of an extensive segment of the Cocos-North America plate boundary during a sequence of Iarge interplate earthquakes that occurred in 8 yr from 1979 to 1986. The sequence included the 1979 Petatlan (Ms = 7.6), the 1981 Playa Azul (Ms = 7.3), 1985 Michoacan (Ms - 8.1), and the 1985 Zihuatanejo (Ms = 7.6) earthquakes, all of which had underthrust faulting mech- anisms with low dip angles and shallow source depths. For this purpose, we have developed complete 3D dynamic rup- ture models for these earthquakes, incorporating their shal- lowly dipping faults rupturing parts of the interface between the subducting and overriding plates, which are embedded in a horizontally layered structure. In these models, some of the kinematic fault parameters previously derived from waveform inversion were used as observational constraints. The spatial distribution of dynamic stress drop and strength excess over the fault have been calculated for each of these earthquakes through an iterative inversion procedure.

It was revealed that heterogeneous stress change over the fault plane of each earthquake came from the failure of a few patchlike, major asperities with medium to high stress drop and from stress increase between and around these as- perities. During the first 1979 Petatlan earthquake, two as- perities in the fault zone located to the southeast of the Mi- choacan segment were ruptured, and two asperities in the central zone of the segment were also ruptured during the 1981 Playa Azul earthquake with a maximum stress drop higher than 80 bars. These ruptures have increased shear stress in the zone between the two events and in the north- western zone of the 1981 event. The 1985 Michoacan earth-

quake ruptured two large-size asperities located at both sides of the 1981 event with maximum stress drops of 80 to 100 bars and two more asperities in the downdip fault section. The stress increase due to this largest event appears to have triggered ruptures on two asperities during the large Zihua- tanejo aftershock in the southeastern adjacent region. A pos- sible explanation for the sequence of events is that the in- crease in stress moves to adjacent regions closer the the critical failure stress and, hence, closer to the time of failure. It was also found that their aftershocks took place mainly in the zones of stress increase outside major asperities, whereas only small numbers of aftershocks occurred within and near the high stress-drop zones. This may be more evidence in- dicating the increase and decrease of stress due to the main- shocks. It should be mentioned that several major asperities that existed on this segment but not overlapped by each other have been ruptured successively so as to fill unbroken gaps on the plate interface. In the Mexican subduction zone, the stress change caused by the previous earthquake seems to have dominant effects on the subsequent events.

Acknowledgments

The present study is essen6alty based on the results of kinematic wave- form inversion made by Carlos Mendoza and Stephan It. Hartzell, to whom we are very ~ateful. Comments given by Jaimson H. Steidl, Arthur F, McGatx (associate editor), and an anonymous reviewer were helpful and improved the manuscript. We also wish to thank Y. Okada for the use of his computer code for static stress drop, Carlos Valdez for providing us aftershock data, and Shri Krishna Singh and Cinna Lomnitz for useful sug- gestions. The senior author (TIVIK.) thanks the financial support provided by the Japan International Cooperation Agency (JICA) for his stay at the

Instituto de Geofisica, UNAM.

Page 17: Dynamic Rupture of Asperities and Stress Change during a ... · Bulletin of the Seismological Society of America, Vol. 88, No. 3, pp. 686-702, June 1998 Dynamic Rupture of Asperities

Dynamic Rupture o f Asperities and Stress Change during a Sequence o f Large Interplate Earthquakes 701

References

Aid, K. (1979). Characterization of barriers on an earthquake fault, J. Geo- phys. Res. 84, 6140-6148.

Anderson, J. G., P. Bodin, J. Brune, J. Prince, S. K. Singh, R. Quaas, M. Onate, and E. Mena (1986). Strong ground motion and source mech- anism of the Mexico earthquake of September 19, 1985, Science 233, 1043-1049.

Astiz, L., H. Kanamori, and H. Eissler (1987). Source characteristics of earthquakes in the Michoacan seismic gap in Mexico, Bull Seism. Soc. Am. 77, 1326-1346.

Beroza, G. and T. Mikumo (1996). Short slip duration in dynamic rupture in the presence of heterogeneous fault properties, J. Geophys. Res. 101, 22449-22460.

Chael, E. P. and G. S. Stewart (1982). Recent large earthquakes along the Middle American Trench and their implications for the subduction process, J. Geophys. Res. 87, 329-338.

DeMets, C., R. G. Gordon, D. F. Argus, and S. Stein (1990). Current plate motions, Geophys. J. Int. 101, 425-478.

Dziewonski, A. M. and J. H. Woodhouse (1983). An experiment in system- atic study of global seismicity: centroid moment tensor solutions for 201 moderate and large earthquakes of 1981, J. Geophys. Res. 88, 3247-3271.

Dziewonski, A. M. and H. H. Woodhouse (1986). Centroid-moment tensor solutions for July-December 1985, Phys. Earth Planet. Interiors 42, 205-214.

Eissler, H., L. Astiz, and K. Kanamori (1986). Tectonic setting and source parameters of the September 19, 1985 Michoacan, Mexico earth- quake, Geophys. Res. Lett. 13, 569-572.

Ekstrom, G. and A. Dziewonski (1986). A very broad band analysis of the Michoacan, Mexico earthquake of September 19, 1985, Geophys. Res. Lett. 13, 605-608.

Fukuyama, E. and T. Mikumo (1993). Dynamic rupture analysis: inversion for the source process of the 1990 Izu-Oshima, Japan earthquake (M6.5), J. Geophys. Res. 98, 6529-6542.

Havskov, J., S. K. Singh, E. Nava, T. Dominguez, and M. Rodriguez (1983). Playa Azul, Michoacan Mexico earthquake of 25 October 1981 (Ms = 7.3), Bull. Seism. Soc. Am. 73, 449-457.

Houston, H. and H. Kanamori (1986). Source characteristics of the 1985 Michoacan, Mexico earthquake at short periods, Geophys. Res. Lett. 13, 597-600.

Ide, S. and M. Takeo (1996). The dynamic rupture process of the 1993 Knshiro-oki earthquake, J. Geophys. Res. 101, 5661-5675.

Kanamori, H. (1981). The nature of seismicity patterns before large earth- quakes, in Earthquake Prediction~An International Review, Maurice Ewing Series 4, D. W. Simpson and P. G. Richards (Editors), Amer- ican Geophysical Union, Washington, D.C., 1-19.

LeFevre, L. V. and K. C. McNally (1985). Stress distribution and subduc- tion of aseismic ridges in the middle America subduction zone, J. Geophys. Res. 90, 4495-4510.

Levander, A. R. (l 985). Use of the telegraphy equation to improve absorb- ing boundary efficiency for fourth-order acoustic wave finite differ- ence scheme, Bull. Seism. Soc. Am. 75, 1847-1852.

McNally, K. C. and J. B. Minster (1981). Nonuniform seismic slip rates along the middle America trench, J. Geophys. Res. 86, 4949-4959.

Mendez, A. J. and J. G. Anderson (1991). The temporal and spatial evo- lution of the 19 September 1985 Michoacan earthquake as inferred from near-source ground-motion records, Bull. Seism. Soc. Am. 81, 844-861.

Mendoza, C. (1993). Coseismic slip of two large Mexican earthquakes from teleseismic body waveforms: implications for asperity interaction in the Michoacan plate boundary segment, J. Geophys. Res. 98, 8197- 8210.

Mendoza, C. (1995). Finite-fault analysis of the 1979 March 14 Petatlan, Mexico, earthquake using teleseismic P waveforms, Geophys. J. lnt. 121, 675-683.

Mendoza, C. and S. Hartzell (1988). Inversion for slip distribution using

teleseismic P waveforms: North Palm Springs, Borah Peak, and Mi- choacan earthquakes, Bull. Seism. Soc. Am. 78, 1092-1111.

Mendoza, C. and S. Hartzell (1989). Slip distribution of the 19 September 1985 Michoacan, Mexico, earthquake: near-source and teleseismic constraints, Bull Seism. Soc. Am. 79, 655-669.

Mikumo, T. and T. Miyatake (1983). Numerical modelling of space and time variations of seismic activity before major earthquakes, Geophys. J. R. Astr. Soc. 14, 559-583.

Mikumo, T. and T. Miyatake (1993). Dynamic rupture processes on a dip- ping fault, and estimates of stress drop and strength excess from the results of waveform inversion, Geophys. J. Int. 112, 481-496.

Mikumo, T. and T. Miyatake (1995). Heterogeneous distribution of dy- namic stress drop and relative fault strength recovered from the results of waveform inversion: 1984 Morgan Hill, California, earthquake, Bull Seism. Soc. Am. 85, 178-193.

Mikumo, T., K. Hirahara, and T. Miyatake (1987). Dynamical fault rnpture processes in heterogeneous media, Tectonophysics 144, 19-36.

Mikumo, T., M. Santoyo, and T. Miyatake (1998). Slip duration in dynamic rupture of large thrust earthquakes (in preparation).

Miyatake, T. (1992). Reconstruction of dynamic rupture process of an earthquake with constraints of kinematic parameters, Geophys. Res. Lett. 19, 349-352.

Okada, Y. (1993). Internal deformation due to shear and tensile faults in a half-space, Bull. Seism. Soc. Am. 82~ 1018-1040.

Pardo, M. and G. Suarez (1995). Shape of the subducted Rivera and Cocos plates in southern Mexico: seismic and tectonic implications, J. Geo- phys. Res. 100~ 12357-12373.

Priestley, K. F. and T. G. Masters (1986). Source mechanism of the Sep- tember 19, 1985 Michoacan earthquake and its implications, Geophys. Res. Lett. 13, 601-604.

Riedesel, M. A., T. H. Jordan, A. F. Sheeman, and P. G. Silver (1986). Moment tensor spectra of the 19 Sept. 85 and 21 Sept. 85 Michoacan, Mexico, earthquake, Geophys. Res. Lett. 13, 609-612.

Ruff, L. J. and A. D. Miller (1994). Rupture process of large earthquakes in the northern Mexico subduction zone, Pageoph 142, 102-171.

Singh, S. K. and F. Mortera ( 1991 ). S ource-time functions of large Mexican subduction earthquakes, morphology of the Benioff zone, age of the plate and their tectonic implications, J. Geophys. Res. 96, 21487- 21502.

Singh, S. K., L. Astiz, and J. Havskov (1981). Seismic gaps and recurrence periods of large earthquakes along the Mexican subduction zone: a re-examination, Bull. Seism. Soc. Am. 71, 827-843.

Singh, S. K., T. Dominguez, R. Castro, and M. Rodriguez (•984). P wave- forms of large shallow earthquakes along the Mexican subduction zone, Bull Seism. Soc. Am, 74, 2135-2156.

Singh, S. K., E. Mena, and R. Castro (1988). Some aspects of source char- acteristics of the 19 September 1985 Michoacan earthquake and ground motion amplification in and near Mexico City from strong motion data, Bull Seism. Soc. Am. 78, 451-477.

Stolte, C., K. C. MeNally, J. Gonzales-Ruiz, G. W. Smith, A. Reyes, C. Rebollar, L. Munguia, and I. Mendoza (1986). Fine structure of a post-failure Wadati-Benioff zone, Geophys. Res. Lett. 13, 577-580.

Suarez, G. and O. Sanchez (1996). Shallow depth of seismogenic coupling in southern Mexico: implications for the maximum size of earth- quakes in the subduction zone, Phys, Earth Planet, hzteriors 93, 53-61.

UNAM Seismology Group (1986). The September 1985 Michoacan earth- quakes: aftershock distribution and history of rupture, Geophys. Res. Lett. 13, 573-576.

Valdez, C. and R. P. Meyer (1996). Seismic structure between the Pacific coast and Mexico City from the Petatlan earthquake (Ms = 7.6) af- tershocks, Geofis. lnt. 35, 377-401.

Valdez, C., R. P. Meyer, and R. Zuniga, J. Havskov, and S. K. Singh (1982). Analysis of the Petatlan aftershocks: numbers, energy release and asperities, J. Geophys. Res. 87, 8519-8527,

Valdez, C. M., W. D. Mooney, S. K. Singh, R. P., Meyer, C. Lorrmitz, J. H. Luetgert, C. E. Helsley, B. T. R. Lewis, and M. Mena 0986).

Page 18: Dynamic Rupture of Asperities and Stress Change during a ... · Bulletin of the Seismological Society of America, Vol. 88, No. 3, pp. 686-702, June 1998 Dynamic Rupture of Asperities

702 T. Mikumo, T. Miyatake, and M. A. Santoyo

Crustal structure of Oaxaca, Mexico, from seismic refraction exper- iments, Bull. Seism. Soc. Am. 76, 547-563.

Yomogida, K. (1988). Crack-like rupture processes observed in near-fault strong motion data, Geophys. Res. Lett. 15, 1223-1226.

Zhang, J. and H. Kanamori (1988). Depth of large earthquakes determined from long-period Rayleigh waves, J. Geophys. Res. 93, 4850-4868.

Instituto de Geofisica Universidad Nacional Autonoma de Mexico Mexico 04510 D.F., Mexico E-mail: mikumo @ oUin.igeofcu.unam.mx

(T.MK., M.A.S.)

Earthquake Research Institute University of Tokyo Tokyo 113, Japan

(T.MY.)

Manuscript received 30 June 1997.