You are on page 1of 10

Journal of South American Earth Sciences 117 (2022) 103885

Contents lists available at ScienceDirect

Journal of South American Earth Sciences


journal homepage: www.elsevier.com/locate/jsames

Seismogenic width in the Guerrero-Oaxaca subduction zone of Mexico


María del Rosario Martínez-López a, *, Carlos Mendoza b, Arturo Iglesias Mendoza a
a
Universidad Nacional Autónoma de México, Instituto de Geofísica, Circuito de La Investigación Cientítifca S/n, Ciudad Universitaria, Delegación Coyoacán, Ciudad de
México, 04510, Mexico
b
Universidad Nacional Autónoma de México, Centro de Geociencias, Campus Juriquilla, Blvd. Juriquilla 3001, Querétaro, 76230, Mexico

A R T I C L E I N F O A B S T R A C T

Keywords: Large earthquakes are generally generated in the contact zone between tectonic plates, with the width of the
Seismicity interplate contact zone controlling the maximum earthquake size. In this work, hypocenters and focal mecha­
Focal mechanisms nisms of earthquakes that occurred in the Guerrero-Oaxaca subduction zone between 1968 and 2017 were
Earthquake rupture zones
examined to determine the maximum depth of seismogenic coupling. From the deepest thrust-faulting events
Seismogenic width
Guerrero-oaxaca subduction zone
consistent with interplate slip, we estimate a seismogenic width of 125 ± 10 km. Aftershock areas of Mw ≥ 6.9
events that occurred from 1937 to 1995 and rupture zones of Mw > 7 earthquakes that occurred from 1995 to
date do not extend beyond a distance of 115–120 km in the downdip direction and are consistent with the
estimated range in seismogenic width. This estimate is greater than the seismogenic width of 60–80 km previ­
ously suggested for the region. The result has important implications for estimating the maximum magnitude
expected for subduction earthquakes and the evaluation of seismic hazard.

1. Introduction (1993) perfomed a study of earthquakes of magnitudes between Mw 6.0


and 7.5 that occurred during the period from 1965 to 1986 in the sub­
The seismogenic zone is defined as the contact between plates that duction zone of Mexico between 95◦ W and 104◦ W longitude. From the
could slip coseismically during large earthquakes at subduction zones analysis of hypocenters and focal mechanisms, they determined that the
(Tichelaar and Ruff, 1993). Interplate events in this contact zone show transition zone from the coupled to the uncoupled zone is between 20
reverse faulting consistent with the geometry of the subducting plate. and 30 km deep. Moreover, Pardo and Suárez (1995), in a study of
The depth of transition from the coupled zone to the uncoupled zone has seismicity and focal mechanisms between 1964 and 1990, proposed a
been determined by identifying the deepest event of thrust faulting maximum coupling depth of 25 km for the Cocos plate in the Mexican
(Comte et al., 1993; Tichelaar and Ruff, 1993). The identification of the subduction zone. From this depth they determined a seismogenic width
maximum depth of the seismogenic zone is necessary to assess the of 60 km (Pardo and Suárez, 1995). Suárez and Sánchez (1996) also
seismic hazard and understand the properties of the faults where meg­ estimated a maximum seismogenic coupling depth of 25 ± 5 km in the
athrust earthquakes can be generated (Li et al., 2015). The maximum subduction region of Michoacan and western Guerrero between 101◦ W
depth of interplate coupling can be determined by identifying the and 103.5◦ W from the analysis of microseismicity and large earth­
deepest events that show thrust faulting with strike approximately quakes. More recently, Martinez-Lopez and Mendoza (2016) suggested a
parallel to the trench and with the orientation of the slip vector deeper seismogenic depth of approximately 40 km in the Michoacan
consistent with the direction of the dip of the subducting plate (Pardo region by examining the transition point between reverse- and
and Suárez, 1995). From the maximum coupling depth, the seismogenic normal-faulting earthquakes. The depth of 40 km obtained for the
width can be determined. Knowledge of the seismogenic width is Michoacan seismogenic zone by Martinez-Lopez and Mendoza (2016) is
important because the maximum earthquake magnitude expected in similar to the maximum depth observed in other subduction zones in the
subduction regions increases with the width (Hyndman et al., 1997). world (Tichelaar and Ruff, 1993; Heuret et al., 2011).
The seismogenic width in the subduction zone of Mexico has been In the Guerrero-Oaxaca subduction zone, the Cocos plate subducts
studied by several researchers (Suárez et al., 1990; Pacheco et al., 1993; below the North American plate with a dip angle of 15◦ (Pardo and
Tichelaar and Ruff, 1993; Pardo and Suárez, 1995). Tichelaar and Ruff Suárez, 1995; Perez-Campos et al., 2008; Kim et al., 2012). Thirteen

* Corresponding author.
E-mail address: rosariomar55@hotmail.com (M.R. Martínez-López).

https://doi.org/10.1016/j.jsames.2022.103885
Received 2 March 2022; Received in revised form 5 June 2022; Accepted 8 June 2022
Available online 18 June 2022
0895-9811/© 2022 Elsevier Ltd. All rights reserved.
M.R. Martínez-López et al. Journal of South American Earth Sciences 117 (2022) 103885

Fig. 1. Epicenters of Mw ≥ 6.9 interplate earthquakes (white stars) that have occurred since 1900 in the Guerrero-Oaxaca subduction zone. The yellow star rep­
resents the possible location of the 1787 event inferred by Suárez et al. (2020). The circles represent the epicenters reported by the National Seismological Service
(SSN) during the period from 2000 to date for events with magnitudes Mw ≥ 4.0 with depth differentiated by the color scale on the right. The yellow rectangle
represents the rupture zone estimated for the 1787 event by Suárez and Albini (2009). The black squares represent the main towns in the area. The thick dashed lines
represent the isodepth contours of the subducting plate (Ferrari et al., 2012). The black line shows the relative convergence rate (Franco et al., 2020). MAT = Middle
America Trench; OGFZ= O. Gorman Fracture Zone.

interplate earthquakes of magnitude greater than or equal to Mw 6.9 more complete examination of the downdip extent of seismogenic
have occurred in this area in the last 120 years (Fig. 1 and Table 1). The coupling in the Guerrero-Oaxaca subduction zone would provide insight
region from Ometepec to Pinotepa Nacional in this subduction zone is into generation of this and other large earthquakes in the region.
considered to be the most active seismic zone of the country There is additional seismicity and focal-mechanism information for
(Núñez-Cornú, 1996). In addition, rupture during the great San Sixto the Guerrero-Oaxaca subduction zone since the results of Pardo and
earthquake of March 28, 1787 appears to have extended northwest into Suárez (1995) were published that can be incorporated into an analysis
the Guerrero-Oaxaca subduction zone (Nuñez-Cornú et al., 2008; of the maximum depth of seismogenic coupling in the region. For
Ramírez-Herrera et al., 2020; Suárez and Albini, 2009; Suárez et al., example, high-quality seismicity catalogs with their respective focal
2020). Suárez et al. (2020) have estimated an intensity magnitude MI of mechanisms have been compiled by Pacheco and Singh (2010) and by
8.6 ± 0.3 for this event from the available macroseismic information. A Yamamoto et al. (2013). In addition, focal mechanisms are available for

2
M.R. Martínez-López et al. Journal of South American Earth Sciences 117 (2022) 103885

Table 1 Project (CMT, 2017) and mechanisms reported from 2002 to 2016 by the
Interplate earthquakes of magnitude greater or equal to 6.9 that have occurred Moment Tensor Project of the National Seismological Service (SSN) of
in the Guerrero-Oaxaca subduction zone (97.2◦ W - 99.7◦ W) since 1900. Mexico (Franco et al., 2020). For the CMT and SSN mechanisms, hy­
Event (day/month/ Latitude Longitude Depth Magnitude pocenters were assigned using locations from the Engdahl-Hilst-Buland
year) (◦ ) (◦ ) (km) (EHB) catalog derived using teleseismic pP, pwP and sP depth phases
15/April/1907 16.62 − 99.221 – Ms 7.79 (Engdahl et al., 1998). CMT mechanisms that were not in the Pardo and
26/March/1908 16.30 − 98.502 – Ms 7.69 Suárez (1995) catalog were also included in the database. Furthermore,
04/Aug/1928 16.83 − 97.613 – Ms 7.49 we reviewed the hypocenters and focal mechanisms obtained by
09/Oct/1928 16.34 97.293 Ms 7.69
− –
Pacheco and Singh (2010) and by Yamamoto et al. (2013) within the
23/Dec/1937 16.39 − 98.613 184 Ms 7.59
14/Dec/1950 16.61 − 98.823 184 Ms 7.19 area of study and added events that were not in the CMT (2017) and
02/Aug/1968 16.01 − 98.013 168, Ms 7.49 Franco et al. (2020) catalogs. Pacheco and Singh (2010) derived
06/June/1982-1 16.40 − 98.345 158 Ms 6.96 mechanisms for events that occurred from 1996 to 2005, and Yamamoto
06/June/1982-2 16.42 − 98.475 208 Ms 7.06 et al. (2013) computed source mechanisms for events that occurred
14/Sept/1995 16.48 98.766 167 Mw 7.37
between 2000 and 2007. We found six mechanisms in the CMT catalog

25/Feb/1996 15.88 − 98.987 157 Mw 7.18
20/March/2012 16.25 − 98.532 202 Mw 7.52 that were duplicated by Yamamoto et al. (2013) and by Pacheco and
16/Feb/2018 16.22 − 98.017 167 Mw 7.28 Singh (2010). The CMT mechanisms, however, were observed to be very
Superscripts indicate references: 1 Nishenko and Singh (1987), 2 UNAM Seis­
similar. Focal mechanisms and hypocenters in the database are given in
mology Group (2013), 3 Santoyo et al. (2005), 4 González-Ruiz and McNally Tables A1 to A5 of the Appendix.
(1988), 5 Astiz and Kanamori (1984), 6 Courboulex et al. (1997), 7 Mendoza and EHB locations obtained using teleseismic phase arrivals correspond
Martínez-López (2021), 8 Kostoglodov and Ponce (1994). to the principal hypocentral information considered in this work. Singh
and Lermo (1985) mention that epicenters in the PDE (Preliminary
events since 1980 from the Global CMT Project (CMT, 2017) and the Determination of Epicenters) of the United States Geological Survey
National Seismological Service (SSN) of Mexico (Franco et al., 2020). In (USGS) and the ISC (International Seismological Centre) bulletins,
this study, we apply the procedure used by Martinez-Lopez and Mendoza which are located using teleseismic data, have a shift of ~35 towards the
(2016) to identify the deepest thrust-faulting events along the NE. Also, in a detailed study of the epicentral differences between the
Guerrero-Oaxaca interplate contact zone to estimate the seismogenic USGS, the SSN, and the Pacheco and Singh (2010) catalogs,
width. In addition, we review Mw ≥ 6.9 earthquakes that have occurred Hjörleifsdóttir et al. (2016) found that, on average, the epicenters re­
in the area in the last 85 years. These events have aftershock areas and ported by global catalogs are shifted by 26 km in the N54◦ E direction
inferred rupture zones that can be examined to identify the downdip compared to those reported by local data. Fig. 2 plots some of the
extent of coseismic rupture to see how it compares with the estimated duplicate events from the EHB catalog with those of Pacheco and Singh
seismogenic width. (2010) and Yamamoto et al. (2013). It is observed that there is an
average difference of ~10 km between the EHB events and those of
2. Data used in the analysis Yamamoto et al. (2013) and Pacheco and Singh (2010). This implies that
the EHB locations may be generally mislocated to the NE by ~10 km.
To identify the maximum depth of seismogenic coupling, a database Depths in the EHB catalog are divided into three categories: L1 with a
was created by combining hypocenters and focal mechanisms from the standard error < 5 km, L2 with an error between 5 and 15 km, and L3
Pardo and Suárez (1995) 1964–1990 catalog with focal mechanisms with an error >15 km (Engdahl et al., 1998). These categories are
reported between 1978 and 2017 by the Global Centroid Moment Tensor indicated in the EHB depths given in the Appendix.

Fig. 2. Comparison of epicenters (circles) of the EHB catalog with those of Pacheco and Singh (2010) and Yamamoto et al. (2013). P&S2010 = Pacheco and Singh
(2010); Yetal 2013 = Yamamoto et al. (2013).

3
M.R. Martínez-López et al. Journal of South American Earth Sciences 117 (2022) 103885

Fig. 3. Epicenters (circles) and focal mechanisms of events examined in this work for the period from 1968 to 2017 in the Guerrero-Oaxaca subduction zone. Red
circles represent reverse faulting consistent with interplate slip. Blue circles indicate normal faulting with orientation and dip consistent with the subducting plate.
Gray circles indicate other types of faulting. Black dashed lines indicate the orientation of the profiles perpendicular to the Middle America Trench (MAT) shown in
Fig. 4. The rectangles represent the width of the profiles. OGFZ= O. Gorman Fracture Zone.

3. Analysis of hypocenter data and focal mechanisms than 70◦ , and a rake between − 50◦ and − 130◦ , and (3) reverse-oblique,
normal-oblique, and other types of earthquakes not in classification 1
We classified the events into three different types: (1) interplate and 2. To visualize the distribution of seismicity at depth, five profiles
thrust earthquakes defined as those events that have a strike of perpendicular to the Middle America Trench (MAT) were drawn in the
240◦ –340◦ , a dip less than 50◦ , and a rake between 30◦ and 130◦ , (2) study region. The direction of the profiles is N15◦ E with a length of 300
normal-faulting earthquakes that have a strike 240◦ –340◦ , a dip less km and a width of 50 km (Fig. 3).

4
M.R. Martínez-López et al. Journal of South American Earth Sciences 117 (2022) 103885

deepest event showing interplate faulting in this profile is event 8 R


located at a depth of 28 km. In the D-D profile (Fig. 4D), the deepest
event showing reverse faulting consistent with interplate slip is event
50C located approximately 105 km from the MAT at a depth of 35 km. In
the E-E ′ profile (Fig. 4E), the deepest event showing reverse faulting is
event 1C at a depth of 28 km at a distance of 105 km away from the
trench This may suggest a smaller seismogenic width in this portion of
the subduction zone, although the event is not significantly deeper than
the interplate thrust events in the other profiles.
To examine the maximum depth of seismogenic coupling in the
Guerrero-Oaxaca subduction zone, profiles AA′ , BB′ , CC’, DD′ and E-E′
were combined into a single profile, plotting only hypocenters of
reverse-faulting events consistent with interplate slip and normal-
faulting events consistent with the downgoing slab (Fig. 5). The deep­
est interplate events in the region are at a depth of 30–35 km. These
events (4C, 50C, 52C) correspond to CMT focal mechanisms with errors
in strike, dip, and rake ≤10◦ estimated from the moment-tensor errors
reported by the global CMT project. The three events have EHB depth
errors ≤5 km. From the maximum depth of the thrust events consistent
with interplate slip and the dip angle, the seismogenic width can be
determined. For a dip angle of 15◦ , which corresponds to an average of
the dip observed at 35 km depth for different plate geometry models (e.
g., Pardo and Suárez, 1995; Dougherty and Clayton, 2014; Hayes et al.,
2018; Manea et al., 2013), we estimate a seismogenic width between
115 and 135 km. Small variations in the assumed dip angle, however,
would result in a greater uncertainty in the seismogenic width. A dif­
ference of ±1◦ in the assumed dip, for example, would yield widths
between 110 and 145 km.

4. Examination of rupture zones and aftershocks of Mw ≥ 6.9


earthquakes

The aftershock areas of the five large earthquakes that occurred from
1937 to 1982 from Table 1 and the rupture zones of the four large events
that occurred since 1995 in the Guerrero-Oaxaca subduction zone were
examined to identify the downdip limit of coseismic rupture. After­
shocks of the events of December 1937 (Ms 7.5) and December 1950 (Ms
7.3) were relocated by González-Ruiz and McNally (1988). For the 1968
event (Mw 7.3), aftershocks were relocated by Tajima and McNally
(1983). For the 1982 earthquake doublet (Mw 6.9; 7.0), aftershocks were
located by Nava (1984). Slip models for the events of September 1995
(Mw 7.3), February 1996 (Mw 7.1), March 2012 (Mw 7.5) and February
2018 (Mw 7.2) were determined by Mendoza and Martínez-López (2021)
by inverting teleseismic P- and SH-waves. A slip model for the March
2012 event was also derived by the UNAM Seismology Group (2013)
using regional and teleseismic data, and the February 2018 event has a
slip model obtained using InSAR and GPS data (Li et al., 2020). In this
Fig. 4. Profiles A-A′ (A), B–B′ (B), C–C′ (C), D-D′ (D) and E-E′ (E) from Fig. 3 for
work, the slip models determined by Mendoza and Martínez-López
the Guerrero-Oaxaca subduction zone. The lateral projection of the focal
(2021) are used to identify the estimated size of the rupture defined by
mechanisms is shown with the labels given in the Appendix. The colors of the
hypocenters and focal mechanisms are as defined in Fig. 3. The black dotted
coseismic slip values greater than zero. These estimated rupture zones
line shows the geometry of the Cocos plate proposed by Hayes et al. (2018). The encompass the source areas observed in other studies of the 2012 and
inverted triangle indicates the position of the trench. 2018 earthquakes.
Fig. 6 shows the events from Table 1 that have reported rupture
Fig. 4 shows the lateral projection of the focal mechanisms onto the zones and aftershock areas. The figure shows aftershock areas of the
five profiles. The hypocenters generally follow the plate geometry of events of magnitude greater than or equal to 6.9 from 1937 to 1982 and
Hayes et al. (2018). In the A-A′ profile (Fig. 4A) the deepest thrust the rupture zones obtained by Mendoza and Martínez-López (2021) for
mechanism (event 73C) is at a depth of 36 and approximately 130 km the 1995, 1996, 2012 and 2018 earthquakes. The 1995 and 2012 rup­
away from the trench with normal faulting events observed at depths tures are shown relative to the epicenters obtained using local data by
greater than 56 km. In profile B–B′ (Fig. 4B) the deepest events that show Courboulex et al. (1997) and UNAM Seismology Group (2013), respec­
interplate faulting are events 9 R and 4C at depths of 32 and 34 km at a tively. For the 1996 and 2018 events, the epicenters are those of the SSN.
distance of 110 and 115 km from the MAT, respectively. The most In western Oaxaca, the rupture zone of the 1996 event extends to the
distant interplate reverse event (52C) is approximately 125 km from the MAT. The rupture zone of the 2018 event is located further inland and
trench and 30 km deep. Event 70 PS in this profile shows extends to approximately 110 km from the MAT. The model determined
normal-faulting within the downgoing slab at a depth of 46 km. Several by Li et al. (2020) for the 2018 event using geodetic data shows a similar
thrust-faulting events are observed in the C–C′ profile (Fig. 4C). The downdip limit. For a dip of 15◦ , this limit would indicate a seismogenic
width of 115 km that is within the range of values estimated from the

5
M.R. Martínez-López et al. Journal of South American Earth Sciences 117 (2022) 103885

Fig. 5. Combination of profiles A-A′ , B–B′ , C–C′ , D-D′ , E-E′ showing hypocenters of events consistent with interplate slip (red circles) and with normal-faulting within
the subducted plate (blue circles). The solid black line shows the plate geometry reported by Hayes et al. (2018). The red line indicates the maximum depth of
interplate seismogenic coupling determined in this study. The inverted triangle indicates the position of the MAT.

Fig. 6. Rupture zones and aftershocks of Mw ≥ 6.9


events that have occurred in the Guerrero-Oaxaca
subduction region since 1937. The polygons repre­
sent the rupture zones of the 1995 (orange), 1996
(blue), 2012 (green) and 2018 (pink) events esti­
mated by Mendoza and Martínez-López (2021). The
ovals represent the aftershock areas of the 1937
(purple), 1950 (brown), 1968 (red) and 1982 (black)
earthquakes. The aftershock zones of the 1937 and
1950 events are from González-Ruiz and McNally
(1988). The aftershock zones of the 1968 and 1982
events are from Tajima and McNally (1983) and Nava
(1984), respectively. The black dashed line represents
the downdip limit of the coseismic rupture zones and
aftershock areas.

seismicity analysis. A review of the variability in slip calculated by analysis of seismicity profiles perpendicular to the MAT, we find that the
Mendoza and Martínez-López (2021) for 400 independent models in­ maximum depth of the seismogenic zone is approximately 30–35 km.
dicates that the rupture areas of the four events generally remain the The results indicate that the seismogenic depth in the Guerrero-Oaxaca
same for variations of ±10◦ in strike, dip and rake. Thus, uncertainties in subduction zone is greater than the 25 km proposed by Pardo and Suárez
the fault parameters used to estimate the rupture areas do not appre­ (1995) for the Cocos plate. The maximum seismogenic depth of 30–35
ciably affect our observation. Also, aftershocks for earthquakes from km is similar to the 40 ± 5 km observed in other subduction margins
1937 to 1982 exhibit a similar downdip limit of 120 km, with the (Tichelaar and Ruff, 1993). The depth is also similar to that observed by
aftershock zone of the 1950 event extending horizontally to approxi­ Martinez-Lopez and Mendoza (2016) in the Michoacan region of the
mately 115 km NE from the MAT. Cocos plate.
Considering a dip angle of 15◦ , a seismogenic interplate width of 125
5. Conclusions and discussion ± 10 km is estimated for the region. Variations of ±1◦ in the assumed
dip angle would indicate widths that vary from 110 to 145 km. This
The hypocenters and focal mechanisms observed between 1968 and suggests that the seismogenic coupling zone is not narrow in the
2017 allowed us to examine the maximum depth of seismogenic Guerrero-Oaxaca subduction zone. According to Herrendörfer et al.
coupling in the Guerrero-Oaxaca subduction zone. From the detailed (2015) a greater seismogenic width favors the generation of a great

6
M.R. Martínez-López et al. Journal of South American Earth Sciences 117 (2022) 103885

earthquake. The seismogenic widths of the 2004 Sumatra-Andaman, earthquakes in the subduction zone. For example, Wyss and Zuñiga
Indonesia (Mw 9.2), 2008 Maule, Chile (Mw 8.8) and 2011 Tohoku, (2016) considered a seismogenic width of 60 km to suggest possible
Japan (Mw 9.0) earthquakes is greater than or equal to 150 km (Tanioka rupture scenarios of maximum magnitudes (M 9) expected in the
et al., 2006; Gulick et al., 2011; Lange et al., 2011; Yagi and Fukahara, Guerrero-Oaxaca subduction zone. The seismogenic width estimated
2011). Our result for the Guerrero-Oaxaca subduction zone suggests that here for the Guerrero-Oaxaca subduction zone would indicate greater
a great earthquake could be generated in the region. Little is known maximum magnitudes for the evaluation of the seismic hazard. The
regarding the downdip width of the 1787 MI 8.6 San Sixto earthquake results also have implications for the rapid derivation of first-order es­
rupture, but it is possible to estimate a width for a rupture length of 450 timates of large-earthquake rupture. Rapid finite-fault inversion pro­
km from current scaling laws. For example, the length (L) in meters to cedures (e.g., Mendoza and Martinez-Lopez, 2022) generally prescribe
width (W) relation of Leonard (2010) for subduction earthqukes. the maximum dimensions of the fault. Therefore, the seismogenic width
determined in this study could be used to define the maximum rupture
log(W) = 0.667log(L) + 1.24 (1)
width allowed for earthquakes in the Guerrero-Oaxaca region to identify
would indicate a rupture width of ~105 km, consistent with the mini­ coseismic slip models useful for early warning and post-earthquake
mum seismogenic width calculated in this study for the Guerrero- response.
Oaxaca region. It should be noted, however, that scaling laws of Leo­
nard (2010) do not consider the 2011 Tohoku, Japan megathrust CRediT authorship contribution statement
earthquake (Mw 9.0).
A review of the aftershock areas and rupture zones of large Mw ≥ 6.9 María del Rosario Martínez-López: Writing – review & editing,
earthquakes that have occurred since 1937 in the region indicates a Writing – original draft, Methodology, Investigation. Carlos Mendoza:
downdip rupture limit of approximately 120 km from the trench, Writing – review & editing, Validation, Supervision, Formal analysis.
consistent with the 125 ± 10 km seismogenic width suggested by the Arturo Iglesias Mendoza: Formal analysis, Conceptualization.
hypocenters and focal mechanisms observed between 1968 and 2017.
Events to the west and southeast of the study region also appear to be Declaration of competing interest
consistent with this width. For example, the rupture area derived by
Ortiz et al. (2000) for the Mw 7.7 event of July 28, 1957 west of the The authors declare that they have no known competing financial
OGFZ exhibits a downdip limit of approximately 120 km. To the interests or personal relationships that could have appeared to influence
southeast, the aftershock zone observed by Singh et al. (1980) for the the work reported in this paper.
event of November 29, 1978 (Ms 7.6) also seems to extend about 100 km
down the dip. Acknowledgments
The seismogenic width estimated in this study is greater than the
value of 60–80 km previously suggested for Cocos-North America plate This Project was financed by a DGAPA-UNAM, Mexico postdoctoral
subduction (e.g., Pardo and Suárez, 1995; Singh et al., 2012). and has grant awarded to M. R. Martinez-Lopez. The images were made with the
implications for the estimation of maximum magnitudes expected for GMT (Generic Mapping Tools) program by Wessel and Smith (1991).

Appendix
Table A1
Hypocenters and focal mechanisms of Pardo and Suárez (1995) analyzed in this study.

Event Date Hypocenter Mechanism Mw

dd/mm/yyyy Latitude (◦ ) Longitude (◦ ) Depth (km) Strike (◦ ) Dip (◦ ) Rake (◦ )

29 PS August 02, 1968 16.273 − 97.976 16.0 287 12 76 6.3


50 PS July 08, 1972 16.015 − 97.167 26.6 278 5 − 90 5.6
70 PS July 18, 1974 16.788 − 98.552 46.3 300 65 − 90 5.5
136 PS June 07, 1982 16.253 − 98.245 20.0 268 10 48 5.8
137 PS June 07, 1982 16.324 − 98.453 11.4 315 24 103 6.0
147 PS February 07, 1983 16.591 − 98.576 25.4 310 17 106 5.5
156 PS June 04, 1984 18.083 − 98.450 65.6 302 48 − 79 5.4
157 PS July 02, 1984 16.592 − 98.638 26.5 302 26 94 5.8
164 PS September 15, 1985 17.795 − 97.335 72.4 309 48 − 85 6.0
184 PS June 07, 1987 16.595 − 98.893 25.0 263 30 58 5.5
185 PS July 15, 1987 17.334 − 97.310 66.1 338 42 − 65 5.8

Table A2
Hypocenters from the EHB bulletin (Engdahl et al., 1998) with Global CMT Project mechanisms analyzed in this study.

Event Date EHB Hypocenter CMT Mechanism Mw

dd/mm/yyyy Latitude (◦ ) Longitude (◦ ) Depth* (km) Strike (◦ ) Dip (◦ ) Rake (◦ )

1C July 31, 1980 16.24 − 97.26 281 277 19 51 5.2


3C February 21, 1983 16.65 − 98.42 291 267 19 50 5.4
4C July 02, 1984 16.77 − 98.48 341 246 36 32 5.9
5C July 04, 1985 17.49 − 97.04 671 318 41 − 98 4.9
7C May 29, 1986 16.93 − 98.80 341 276 59 81 5.2
10C April 25, 1989 16.77 − 99.27 241 276 10 66 6.2
11C April 01, 1991 16.15 − 98.29 191 40 51 40 5.4
(continued on next page)

7
M.R. Martínez-López et al. Journal of South American Earth Sciences 117 (2022) 103885

Table A2 (continued )
Event Date EHB Hypocenter CMT Mechanism Mw

dd/mm/yyyy Latitude (◦ ) Longitude (◦ ) Depth* (km) Strike (◦ ) Dip (◦ ) Rake (◦ )

12C April 01, 1991 16.13 − 98.21 253 55 39 58 5.1


13C April 07, 1991 16.33 − 97.55 412 224 63 43 5.1
14C November 24, 1991 16.35 − 97.95 261 278 36 96 5.2
15C June 07, 1992 16.50 − 98.72 152 241 16 48 5.3
16C June 07, 1992 16.38 − 98.70 152 278 19 88 5.2
17C May 15, 1993 16.70 − 98.34 241 242 35 30 6.0
18C August 05, 1993 17.38 − 98.33 541 294 39 − 135 4.8
19C November 13, 1993 16.27 − 98.71 131 262 16 66 5.5
20C December 13, 1994 16.25 − 98.46 181 270 14 67 5.3
21C September 14, 1995 16.83 − 98.63 231 289 15 85 6.3
22C October 30, 1995 16.44 − 98.41 312 262 17 50 4.9
23C February 25, 1996 15.92 − 98.12 241 280 16 74 6.0
24C February 25, 1996 16.05 − 97.97 153 253 16 56 5.5
25C February 26, 1996 15.82 − 97.87 173 347 11 92 5.0
26C March 13, 1996 16.65 − 98.95 221 128 29 100 5.1
27C March 19, 1996 15.80 − 97.33 141 224 17 22 5.7
28C March 20, 1996 15.73 − 97.36 181 283 13 60 5.3
29C March 20, 1996 15.88 − 97.29 153 246 16 47 5.1
30C March 27, 1996 16.47 − 98.09 26 266 21 70 5.5
31C January 21, 1997 16.42 − 98.06 273 281 26 49 5.0
32C July 19, 1997 16.19 − 98.21 201 282 14 78 5.5
33C December 16, 1997 16.17 − 98.87 191 260 18 64 5.3
34C March 05, 1998 16.14 − 98.31 212 269 16 79 5.2
35C April 04, 1999 15.80 − 97.48 141 276 20 82 5.0
36C May 05, 1999 14.30 − 94.79 111 313 44 − 73 5.9
37C November 10, 2001 16.20 − 98.16 133 243 10 69 5.1
38C June 19, 2002 16.29 − 97.94 231 272 11 60 5.0
39C August 27, 2002 16.16 − 97.37 261 233 58 1 5.0
40C November 15, 2004 16.34 − 98.42 153 295 21 94 5.1
41C November 15, 2004 16.10 − 98.44 222 306 26 110 4.9
42C November 10, 2006 15.73 − 97.25 132 310 40 102 4.8
43C March 15, 2007 16.19 − 97.04 241 63 8 35 4.9
44C March 30, 2007 16.22 − 98.82 153 287 19 75 5.0
45C May 04, 2007 17.29 − 96.71 591 176 46 − 46 5.0
46C August 26, 2007 16.31 − 97.91 241 288 22 90 4.8
47C May 17, 2008 16.35 − 97.89 251 280 18 68 4.8
48C April 16, 2010 16.34 − 98.26 152 276 6 90 4.8
49C April 20, 2010 16.35 − 98.23 232 187 46 58 4.9
50C June 30, 2010 16.42 − 97.79 351 286 12 72 5.8
51C May 05, 2011 16.85 − 98.64 241 54 12 22 5.3
52C March 21, 2012 16.79 − 98.25 301 273 22 67 5.0
53C March 24, 2012 16.45 − 98.08 201 274 25 66 4.8
54C April 01, 2012 16.68 − 98.31 281 293 26 81 5.2
55C April 02, 2012 16.27 − 98.58 152 98 44 72 5.0
56C April 13, 2012 16.40 − 98.12 141 44 28 44 5.4
57C April 13, 2012 16.41 − 98.03 121 98 26 132 4.7
58C July 24, 2012 16.51 − 98.07 191 273 16 52 5.2
59C September 09, 2012 16.05 − 98.16 222 270 34 79 4.7
60C September 22, 2012 16.52 − 98.05 202 264 24 58 5.3
61C November 21, 2012 16.32 − 98.38 142 299 28 97 4.6
62C March 26, 2013 16.18 − 98.19 121 273 21 59 5.5
63C March 26, 2013 16.22 − 98.20 162 274 17 67 5.1
64C January 02, 2014 16.13 − 97.82 232 120 48 116 4.8
65C March 10, 2014 16.03 − 98.33 203 276 26 76 5.6
66C March 13, 2014 16.02 − 98.62 142 291 22 95 4.8
67C May 24, 2014 16.52 − 98.17 162 279 13 68 5.4
68C August 13, 2014 16.29 − 98.21 101 252 25 − 43 5.5
69C September 08, 2014 16.29 − 98.22 182 275 48 − 39 4.7
70C January 12, 2015 16.34 − 97.91 252 327 34 111 4.9
71C April 05, 2015 16.37 − 98.49 152 294 27 97 4.9
72C June 25, 2015 16.69 − 97.97 251 238 40 − 4 5.0
73C November 23, 2015 17.03 − 98.79 361 290 61 111 5.5
74C March 24, 2016 16.33 − 97.89 211 290 19 71 4.9
75C May 08, 2016 16.40 − 97.74 201 294 15 75 5.9
76C June 27, 2016 16.40 − 97.84 201 293 16 83 5.7
77C July 19, 2016 17.57 − 98.68 562 312 37 − 98 5.0
78C August 11, 2016 16.33 − 99.05 153 296 28 98 4.8
79C December 21, 2017 16.13 − 98.28 131 278 24 70 5.1
*
Subscripts indicate depth error: 1 ≤5 km, 2 5–15 km and 3 > 15 km.

8
M.R. Martínez-López et al. Journal of South American Earth Sciences 117 (2022) 103885

Table A3
Hypocenters from the EHB bulletin (Engdahl et al., 1998) with focal mechanisms of the SSN (Franco et al., 2020) analyzed in this study.

Event Date EHB Hypocenter SSN Mechanism Mw

dd/mm/yyyy Latitude (◦ ) Longitude (◦ ) Depth* (km) Strike (◦ ) Dip (◦ ) Rake (◦ )

1R May 28, 2002 16.45 − 99.37 103 108 48 104 4.8


2R January 03, 2007 16.16 − 97.34 261 121 50 89 4.3
3R March 01, 2007 16.57 − 98.89 251 109 73 72 4.5
4R December 09, 2008 17.98 − 98.32 551 115 54 − 102 4.5
5R February 03, 2010 17.99 − 98.12 561 313 48 − 66 4.8
6R February 14, 2010 16.21 − 98.55 153 297 47 94 4.2
7R December 16, 2011 16.34 − 98.07 272 288 46 92 4.7
8R January 15, 2013 16.65 − 98.27 282 130 78 105 4.6
9R August 13, 2013 16.68 − 98.37 321 290 18 81 5.2
10 R November 28, 2015 17.25 − 96.66 60 147 47 − 81 4.5
11 R July 24, 2016 17.97 − 97.31 642 119 78 − 98 4.4
12 R July 26, 2016 15.93 − 98.03 153 114 89 112 4.8
*
Subscripts indicate depth error: 1 ≤5 km, 2 5–15 km and 3 > 15 km.

Table A4
Hypocenters and focal mechanisms of Pacheco and Singh (2010) analyzed in this study.

Event Date Hypocenter Mechanism Mw

dd/mm/yy Latitude ( )

Longitude ( )◦
Depth (km) Strike ( )

Dip ( )

Rake ( )◦

10 P August 27, 1996 16.659 − 99.320 21.8 250 42 55 4.2


41 P September 30, 1998 16.456 − 99.073 16.6 340 62 120 4.1
44 P February 15, 1999 16.541 − 99.308 27.3 79 71 − 37 4.2
50 P August 13, 1999 16.614 − 99.208 19.6 236 49 35 3.9
60 P March 21, 2000 16.570 − 99.041 16.7 269 34 66 4.8
96 P August 29, 2002 16.813 − 98.911 28.8 274 46 90 4.0
116 P June 14, 2004 16.200 − 98.139 17.4 277 11 70 5.9
125 P December 02, 2005 16.378 − 98.489 27.6 309 27 96 4.8

Table A5
Hypocenters and focal mechanisms of Yamamoto et al. (2013) analyzed in this study.

Event Date Hypocenter Mechanism Mw

dd/mm/yy Latitude (◦ ) Longitude (◦ ) Depth (km) Strike (◦ ) Dip (◦ ) Rake (◦ )

1Y March 14, 2000 16.14 − 98.47 20 117 63 75 4.2


2Y March 29, 2000 16.10 − 98.78 16 110 52 93 3.8
6Y January 06, 2004 16.22 − 98.64 53 292 68 98 4.0
7Y March 02, 2004 16.56 − 98.50 15 114 48 89 4.3
11Y May 19, 2005 16.54 − 98.54 48 184 80 152 4.8
13Y November 07, 2005 16.63 − 98.46 43 326 52 87 4.5
14Y March 25, 2006 16.02 − 98.97 26 95 53 64 4.6
15Y December 05, 2006 16.11 − 98.69 20 117 54 90 4.6
16Y June 12, 2007 16.31 − 98.52 25 67 76 84 4.6
17Y October 02, 2007 16.24 − 98.85 25 318 48 − 88 4.1
18Y November 08, 2007 16.56 − 98.45 20 312 74 − 77 3.9
19Y December 07, 2007 16.59 − 98.41 29 261 76 79 4.5
21Y May 08, 2009 16.58 − 98.45 16 223 84 − 82 5.2
23Y April 21, 2001 15.63 − 98.31 26 109 53 69 4.3
24Y May 20, 2001 16.32 − 97.98 18 147 86 131 4.1
25Y July 21, 2001 16.17 − 98.20 26 272 81 − 87 3.7
29Y July 11, 2002 16.36 − 98.03 5 302 86 − 155 4.3
31Y November 17, 2003 16.31 − 97.93 16 94 48 − 91 4.0
35Y February 14, 2009 16.03 − 97.84 17 346 85 − 171 4.0
36Y May 10, 2000 16.06 − 97.26 25 120 64 102 4.1
37Y January 13, 2004 16.02 − 97.17 25 322 47 103 5.1
38Y January 13, 2004 16.16 − 97.22 21 304 80 93 4.2
39Y January 14, 2004 16.06 − 97.18 36 304 80 93 4.6
40Y February 01, 2004 16.17 − 97.26 59 74 69 32 4.4
42Y September 28, 2005 16.03 − 97.24 33 329 57 − 42 3.9
44Y August 19, 2006 16.04 − 97.42 46 287 43 84 4.2
46Y February 25, 2007 16.19 − 97.26 19 125 72 100 3.6
49Y April 01, 2007 16.25 − 97.34 25 293 87 91 4.4

9
M.R. Martínez-López et al. Journal of South American Earth Sciences 117 (2022) 103885

References Nava, E., 1984. Estudio de los temblores de Ometepec del 7 de junio de 1982, y sus
réplicas. Tesis Licenciatura (Ingeniero Geofísico). Facultad de Ingeniería, UNAM.
Nishenko, S.P., Singh, S.K., 1987. The Acapulco-Ometepec, Mexico, earthquakes of 1907-
Astiz, L., Kanamori, H., 1984. An earthquake doublet in Ometepec, Guerrero, Mexico.
1982: evidence for a variable recurrence history. Bull. Seismol. Soc. Am. 77,
Phys. Earth Planet. In. 34, 24–45.
1359–1367.
Cmt, 2017. Global CMT Project. available in. http://www.globalcmt.org/, 2019, 4.
Núñez-Cornú, F.J., 1996. A doublet seismic front and earthquake cycles along the coast
Courboulex, F., Singh, S.K., Pacheco, J.F., 1997. The 1995 Colima-Jalisco, Mexico,
of Oaxaca, Mexico. Seismol Res. Lett. 67 (6), 33–39.
earthquake (Mw = 8): a study of the rupture process. Geophys. Res. Lett. 24,
Nuñez-Cornú, F.J., Ortiz, M., Sanchez, J.J., 2008. The great 1787 Mexican tsunami. Nat.
1019–1022.
Hazards 47, 569–576. https://doi.org/10.1007/s11069-008-9239-1.
Dougherty, S.L., Clayton, R.W., 2014. Seismicity and structure in central Mexico:
Ortiz, M., Singh, S.K., Kostoglodov, V., Pacheco, J., 2000. Source areas of the Acapulco-
evidence for a possible slab tear in the South Cocos plate. J. Geophys. Res. Solid
San Marcos, Mexico earthquakes of 1962 (M 7.1; 7.0) and 1957 (M 7.7), as
Earth 119, 3424–3447.
constrained by tsunami and uplift records. Geofisc. Int. 39 (4), 337–348.
Engdahl, E.R., Hilst, V., Buland, R.D., 1998. Global teleseismic earthquake relocation
Pacheco, J.F., Singh, S.K., 2010. Seismicity and state of stress in Guerrero segment of the
with improved travel times and procedures for depth determination. Bull. Seismol.
Mexican subduction zone. J. Geophys. Res. 115, B01303 https://doi.org/10.1029/
Soc. Am. 88, 722–743.
2009JB006453.
Franco, S.I., Iglesias, A., Fukuyama, E., 2020. Moment tensor catalog for Mexican
Pacheco, J.F., Sykes, L.R., Scholz, Ch H., 1993. Nature of seismic coupling along simple
earthquakes: almost two decades of seismology. Geofisc. Int. 59 (2), 54–82.
plate boundaries of the subduction type. J. Geophys. Res. 98 (B8), 14,133–14,159.
González-Ruiz, J.R., McNally, K.C., 1988. Stress accumulation and release since 1882 in
Pardo, M., Suárez, G., 1995. Shape of the subducted Rivera and Cocos plates in southern
Ometepec, Guerrero, Mexico: implications for failure mechanisms and risk
Mexico: seismic and tectonic implications. J. Geophys. Res. 100 (B7), 12357–12373.
assessments of a seismic gap. J. Geophys. Res. Solid Earth 93 (B6), 6297–6317.
Perez-Campos, X., Kin, Y., Husker, A., Davis, P.M., Clayton, R.W., Iglesias, A., Pacheco, J.
Gulick, S.P.S., James, A.A., McNeil, L., Nathan, L.B., Martin, K.M., Henstock, T., Bull, J.
F., Singh, S.K., Manea, V.C., Gurnis, M., 2008. Horizontal subduction and truncation
M., Dean, S., Djajadihardja, Y.S., Permana, H., 2011. Updip rupture of the 2004
of the Cocos plate beneath central Mexico. Geophys. Res. Lett. 35 (18), l18303.
Sumatra earthquake extended by thick indurated sediments. Nat. Geosci. 4,
Ramírez-Herrera, M.T., Corona, N., Cerny, J., 2020. Sand deposits reveal great
453–456. https://doi.org/10.1038/ngeo1176.
earthquakes and tsunamis at Mexican Pacific coast. Sci. Rep. 10, 11452 https://doi.
Hayes, G.P., Moore, G.L., Portner, D.E., Hearne, M., Flamme, H., Furtney, M., Smoczk, G.
org/10.1038/s41598-020-68237-2.
M., 2018. Slab2, a comprehensive subduction zone geometry model. Geophysics,
Santoyo, M.A., Singh, S.K., Mikumo, T., Ordaz, M., 2005. Space-time clustering of large
Science 362, 58–61.
thrust earthquakes along the Mexican subduction zone: an evidence of source stress.
Herrendörfer, R., Dinther, V., Gerya, Y., 2015. Earthquake supercycle in subduction
Bull. Seismol. Soc. Am. 95 (5), 1856–1864.
zones controlled by the width of the seismogenic zone. Nat. Geosci. 8 (6), 471–474.
Seismology Group, U.N.A.M., 2013. Ometepec-pinotepa nacional, Mexico earthquake of
https://doi.org/10.1038/ngeo2427.
20 March 2012 (Mw 7.5): a preliminary report. Geofisc. Int. 52 (2), 173–196.
Heuret, A., Lallemand, S., Funiciello, F., Piromallo, C., Faccenna, C., 2011. Physical
https://doi.org/10.1016/S0016-7169(13)71471-5.
characteristics of subduction interface type seismogenic zones revisited. G-cubed 12,
Singh, S.K., Lermo, J., 1985. Mislocations of Mexican earthquakes as reported in
Q01004. https://doi.org/10.1029/2010GC003230.
international bulletins. Geofisc. Int. 24 (2), 333–351.
Hjörleifsdóttir, V., Singh, S.K., Husker, A., 2016. Differences in epicentral location of
Singh, S.K., Havskov, J., McNally, K., Ponce, L., Hearn, T., 1980. The Oaxaca Mexico,
Mexican earthquakes between local and global catalogs: an update. Geofisc. Int. 55
earthquake of 29 de November 1978: 1 Preliminary report on aftershocks. Geofisc.
(1), 79–93.
Int. 335–340.
Hyndman, R.D., Yamano, M., Oleskevich, D.A., 1997. The seismogenic zone of
Singh, S.K., Pérez-Campos, X., Iglesias, A., Melgar, D., 2012. A method for rapid
subduction thrust faults. Isl. Arc 6 (3), 244–260.
estimation of moment magnitude for early tsunami warning based on coastal GPS
Kim, Y.H., Miller, M.S., Pearce, F., Clayton, R.W., 2012. Seismic imaging of the Cocos
networks. Seismol Res. Lett. 83 (3), 516–530.
plate subduction zone system in central Mexico. G-cubed 13 (7), 1–16.
Suárez, G., Albini, P., 2009. Evidence for great tsunamigenic earthquakes (M 8.6) along
Kostoglodov, V., Ponce, L., 1994. Relationship between subduction and seismicity in the
the Mexican subduction zone. Bull. Seismol. Soc. Am. 99 (2A), 892–896.
Mexican part of the Middle America trench. J. Geophys. Res. 99 (B1), 729–742.
Suárez, G., Sánchez, O., 1996. Shallow depth of seismogenic coupling in southern
Lange, D., Tilmann, F., Barrientos, S.E., Contreras-Reyes, E., Methe, P., Moreno, M.,
Mexico: implications for the maximum size of earthquakes in the subduction zone.
Heit, B., Agurto, H., Bernard, P., Vilotte, J.P., 2011. Aftershock seismicity of the 27
Phys. Earth Planet. In. 93, 53–61.
February 2010 Mw 8.8 Maule earthquake rupture zone. Earth Planet Sci. Lett.
Suárez, G., Monfret, T., Wittlinger, G., David, C., 1990. Geometry of subduction and
317–318, 413–425. https://doi.org/10.1016/j.epsl.2011.11.034.
depth of the seismogenic zone in the Guerrero Gap, México. Nature 345, 336–338.
Leonard, M., 2010. Earthquake fault scaling: self -consistent relation of rupture length,
Suárez, G., Ruiz-Barón, D., Chico-Hernández, C., Zúñiga, F.R., 2020. Catalog of
width, average displacement, and moment release. Bull. Seismol. Soc. Am. 100 5A,
preinstrumental earthquakes in central Mexico: epicentral and magnitude
1971–1988.
estimations based on macroseismic data. Bull. Seismol. Soc. Am. 110 (6),
Li, J., Shillington, D.J., Becel, A., Nedimovic, M.R., Webb, S.C., Saffer, D.M., Keranen, K.
3021–3036. https://doi.org/10.1785/0120200127.
M., Kuehn, H., 2015. Downdip variations in seismic reflection character:
Tajima, F., McNally, K.C., 1983. Seismic rupture patterns in Oaxaca, Mexico. J. Geophys.
implications for fault structure and seismogenic behavior in the Alaska subduction
Res. 88 (B5), 4263–4275.
zone. J. Geophys. Res. Solid Earth 120, 7883–7904. https://doi.org/10.1002/
Tanioka, Y., Kususose, T., Kathiroli, S., Nishimura, Y., Iwasaki, S.I., Satake, K., 2006.
2015jb012338.
Rupture process of the 2004 great Sumatra-Andaman earthquake estimated from
Li, Y., Shan, X., Zhu, Ch, Qiao, X., Zhao, L., Qu, Ch, 2020. Geodetic model of the 2018
tsunami waveforms. Earth Planets Space 58, 203–209.
Mw 7.2 Pinotepa, Mexico, earthquake inferred from InSAR and GPS Data. Bull.
Tichelaar, B.W., Ruff, L.J., 1993. Depth of seismic coupling along subduction zones.
Seismol. Soc. Am. 110 (3), 1115–1124.
J. Geophys. Res. Solid Earth 98 (B2), 2017–2037.
Manea, V.C., Manea, M., Ferrari, L., 2013. A geodynamical perspective on the subduction
Wessel, P., Smith, W.H.F., 1991. Free software helps map and display data. Eos
of Cocos and Rivera plates beneath Mexico and Central America. Tectonophysics
Transactions: American Geophysical Union 72 (41), 441–448.
609, 56–81.
Wyss, M., Zuñiga, F.R., 2016. Estimated casualities in a possible great earthquake along
Martinez-Lopez, M.R., Mendoza, C., 2016. Acoplamiento sismogénico en la zona de
the Pacific coast of Mexico. Bull. Seismol. Soc. Am. 106 (4), 1867–1874.
subducción de Michoacán-Colima-Jalisco, Mexico. Bol. Soc. Geol. Mex. 68 (2),
Yagi, Y., Fukahara, Y., 2011. Rupture process of the 2011 Tohoku-oki earthquakes and
199–214.
absolute elastic strain release. Geophys. Res. Lett. 38, L19307.
Mendoza, C., Martinez-Lopez, M.R., 2022. Rapid finite-fault analysis of large Mexico
Yamamoto, J., González-Moran, T., Quintanar, L., Zavaleta, A.B., Zamora, A.,
earthquakes using teleseismic P waves. J. Seismol. 26, 333–342. https://doi.org/
Espindola, V.H., 2013. Seismic patterns of the Guerrero-Oaxaca, Mexico region, and
10.1007/s10950_022_1083_y online.
its relationship to the continental margin structure. Geophys. J. Int. 192, 375–389.
Mendoza, C., Martínez-López, M.R., 2021. Rupture models of recent Mw>7 thrust
Comte, D., Pardo, M., Dorbath, L., Dorbath, C., Haessler, H., Rivera, L., Cisternas, A.,
earthquakes in the Guerrero–Oaxaca region of the Mexico subduction zone using
Ponce, L., 1993. Seismogenic interplate contact zone and crustal seismicity around
teleseismic body waves. Seismol Res. Lett. 92 (6), 3565–3576. https://doi.org/
Antofagasta, northern Chile using local data. Geophysical Journal International
10.1785/0220200423.
submit.

10

You might also like