2019 Ammirati Et Al PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Bulletin of the Seismological Society of America, Vol. 109, No. 5, pp. 1985–1999, October 2019, doi: 10.

1785/0120190082

The Crustal Seismicity of the Western Andean Thrust (Central


Chile, 33°–34° S): Implications for Regional Tectonics
and Seismic Hazard in the Santiago Area
by Jean-Baptiste Ammirati, Gabriel Vargas, Sofía Rebolledo, Rachel Abrahami,
Bertrand Potin, Felipe Leyton, and Sergio Ruiz

Abstract Most of the recorded seismicity in central Chile can be linked to the sub-
duction of the Nazca plate. To the east, a much smaller fraction is observed at 0–30 km
depths beneath the western Andean thrust. Paleoseismic studies evidenced the occur-
rence of at least two major earthquakes (M > 7) over the past 17 ka, associated with the
San Ramón fault (SRF): an important tectonic feature characterizing the west Andean
thrust, close the Santiago metropolitan area. To better constrain the crustal seismicity in
this area, the Chilean Seismological Center (CSN) extended its permanent seismic net-
work with seven new broadband seismometers deployed around the scarp of the SRF
and farther east. The improved azimuthal distribution and reduced station spacing
allowed to complete the CSN catalog with more than 900 smaller magnitude earth-
quakes (M L < 2:5) detected and located within the study region. The use of a 3D veloc-
ity model derived from P- and S-wave travel-time tomography considerably lowered the
uncertainties associated with hypocentral locations. Our results show an important seis-
micity beneath the Principal Cordillera located at a depth of ∼10 km, and a deeper
seismicity (~15 km) aligned with the main Andean thrust more to the west, parallel
to the scarp of the SRF. Regional stress inversion results suggest that the seismicity
of the west Andean thrust is accommodating northeast–southwest compressional stress,
consistent with the convergence of the Nazca plate. Based on our improved crustal seis-
micity, combined with observations from previous studies, we have been able to refine
the scenario of an Mw 7.5 earthquake rupturing the SRF. Ground-motion prediction
results show peak ground accelerations of ∼0:8g close to the fault scarp.

Supplemental Content: Cross sections of the 3D velocity model used in this


study, probabilistic power spectral density (PPSD) plots for the stations of the
Chilean Seismological Center (CSN) network located in the study region as well the
waveforms corresponding to a seismic event that occurred beneath the west Andean
thrust, and tables of hypocentral location and associated errors of the 917 events char-
acterized in this study, and the focal mechanism solutions determined from P-wave
first-motion polarity.

Introduction and Geological Setting


Central Chile marks the boundary between the Nazca Valdivia earthquake or more recently, the 2010 Mw 8.9
plate and the South American plate, which makes it one of Maule earthquake (Cifuentes and Silver, 1989; Vigny et al.,
the most seismically active countries on the planet. In this 2011; Ruiz and Madariaga, 2018) as well as intermediate
region (Fig. 1), roughly 95% of the recorded seismicity depth intraplate events, mainly extensional, that have been
can be linked to the subduction of the Nazca plate under related to dehydration and brittle failure within the sub-
the South American plate, along the Chilean margin. This ducting Nazca plate (Beck et al., 1998; Barrientos et al.,
major seismogenic zone is characterized by the occurrence 2004; Ruiz et al., 2019). Around 5% of the seismicity
between 0 and 50 km depths of some of the largest thrust recorded in central Chile can be observed between 0 and
earthquakes ever recorded such as the 1960 M w 9.5 30 km depths beneath the western flank of the Andean

1985

Downloaded from https://pubs.geoscienceworld.org/ssa/bssa/article-pdf/109/5/1985/4833931/bssa-2019082.1.pdf


by Univ de Chile user
1986 J.-B. Ammirati, G. Vargas, S. Rebolledo, R. Abrahami, B. Potin, F. Leyton, and S. Ruiz

Shin and Teng, 2001; Alvarado et al., 2005; Quigley et al.,


2016; Zhang et al., 2016; Instituto Nacional de Prevención
Sísmica [INPRES], 2019). To this day, the 1958 M w 6.3 Las
Melosas earthquake is the strongest crustal event to have
occurred in our study region that was instrumentally
recorded and characterized (Pardo and Acevedo, 1984;
Alvarado et al., 2009). It produced several casualties, build-
ing collapses, and severely damaged the water distribution
infrastructure of the Santiago area (Sepúlveda et al., 2008).
The city of Santiago de Chile, located in the central valley
on the western Andean piedmont, has expanded over the sedi-
ments of the Central Depression, mainly Quaternary detrital
material carried from the Mapocho and the Maipo rivers
(Fig. 2). An important topographic scarp (∼2700 m) directly to
the east marks the beginning of the Principal Cordillera (PC)
mostly constituted by Cenozoic volcanic rocks of the Abanico
and Farellones formations, intrusive, and highly deformed
Jurassic to Cretaceous sequences (Thiele, 1980). Farther east,
the Frontal Cordillera (FC) exhibits Permian–Triassic volcanic
rocks and intrusive granitoids together with late Paleozoic
sequences and Proterozoic metamorphic rocks (Fig. 2).
Geological observations in this regions showed that the
present structural configuration of the Andes between 33°
and 35° S is the result of intense crustal shortening within
the upper plate, in the context of subduction (Allmendinger
et al., 1990; Cristallini and Ramos, 2000; Giambiagi and
Ramos, 2002). It appears that most of this crustal shortening
was accommodated since ∼16 Ma by east-vergent structures
on the eastern side (Giambiagi, Ramos, et al., 2003). In the
western side, the contact between the Abanico formation and
the Quaternary sediments of the Central Depression is defined
by a west-verging reverse fault that accommodated part of the
crustal shortening in the western side: the San Ramón fault
(SRF; Armijo et al., 2010; Vargas et al., 2014). Although
Figure 1. Seismicity of central Chile recorded by the Chilean the extension in depth of the aforementioned structures and
National Network (CSN) between 2000 and 2017. (Inset) Location
of our study region at the western margin of South America. Stars their tectonic implications are still debated (Armijo et al.,
show the epicenter of recent megathrust earthquakes along the 2010; Giambiagi et al., 2015; Riesner et al., 2018), the
Chilean margin. Squares mark the location of the CSN permanent SRF certainly raises questions in terms of seismic hazard
seismic stations. Note the strike-slip mechanism associated with because of its proximity with the city of Santiago (Fig. 2).
three crustal events from the Global Centroid Moment Tensor
Recent paleoseismic studies (Vargas et al., 2014) evidenced
(CMT) catalog (Ekström et al., 2012) and Alvarado et al.
(2005). The solid arrow shows the regional velocity field direction two major seismic events occurred on the SRF, respectively
inferred from 20 yr of Global Navigation Satellite Systems (GNSS) ∼8 and ∼17 ka. The slip associated with these events has been
measurements (Métois et al., 2016). The color version of this figure estimated to 4.7–4.9 m corresponding to seismic events of
is available only in the electronic edition. magnitude 7:2 < M w < 7:4. In the metropolitan area of
Santiago (0–20 km away from the SRF scarp), such magni-
Cordillera (Fig. 1). Although the occurrence of seismic tudes would generate ground accelerations much stronger than
events in this sector is far less frequent than the Chilean mar- the ones observed during high-magnitude megathrust earth-
gin, near-surface rupturing represents a considerable threat to quakes, along the Chilean margin (>100 km west of the city).
urban centers located in this region. The Santiago metropoli- As an example, the 2010 M w 8.8 Maule earthquake produced
tan area is home to nearly half of the total Chilean population a peak ground acceleration (PGA) of ∼0:3g in the Santiago
and concentrates an important fraction of the Chilean eco- area (Boroschek et al., 2012). Previous seismic hazard studies
nomic activity. There is no lack of examples to illustrate the simulated ground accelerations generated by rupturing the
threat of crustal faults worldwide: Mendoza (Argentina) SRF (Pérez et al., 2014; Estay et al., 2016) and estimated
1862; San Juan (Argentina) 1944; Kobe (Japan) 1995; PGAs of ∼0:7g near the fault scarp.
Chi-Chi (Taiwan) 1999; Christchurch (New Zealand) 2010; In this work, we use continuous seismic waveforms
and Gorkha (Nepal) 2015 (Kanaori and Kawakami, 1996; recorded by the Chilean seismic network to detect and

Downloaded from https://pubs.geoscienceworld.org/ssa/bssa/article-pdf/109/5/1985/4833931/bssa-2019082.1.pdf


by Univ de Chile user
The Crustal Seismicity of the Western Andean Thrust (Central Chile, 33°–34° S) 1987

Figure 2. First-order geology of the western margin of South America around 32° S and morphostructural map of the study region.
Geological information has been compiled from Armijo et al. (2010), Farías et al. (2010), Giambiagi et al. (2015), and Riesner et al. (2018).
The gray polygon delimits the metropolitan area of Santiago. AFTB, Aconcagua fold and thrust belt; SRF, San Ramón fault; WAFTB, west
Andean fold and thrust belt. Note the change in structure vergence from east to west, in the Principal Cordillera (PC). The color version of this
figure is available only in the electronic edition.

precisely locate the low-magnitude (ML < 2:5) seismicity Previous Seismological Studies and Current
beneath the west Andean thrust. We also use the first-motion Operations
polarity information to determine the focal mechanism asso-
The Chilean Seismological Center (Centro Sismológico
ciated with some of these events to estimate the regional
Nacional or CSN) of the University of Chile started to oper-
stress principal directions. This information, along with the ate in March 2013 as a continuation of the Seismological
seismic location allows us to discuss about the differences Service of the Department of Geophysics of the same
and similarities between our results, tectonic models, and university, which was in charge of the permanent seismic
interpretations made in previous studies. In addition, we operations since 1982.
use our seismic results to refine the scenario of near-surface To better comprehend the crustal seismicity beneath the
rupture in the Santiago area. We then calculate the expected PC, Barrientos et al. (2004) relocated the events recorded by
ground accelerations for the metropolitan area of Santiago. the CSN between 1986 and 2001 using an improved 1D

Downloaded from https://pubs.geoscienceworld.org/ssa/bssa/article-pdf/109/5/1985/4833931/bssa-2019082.1.pdf


by Univ de Chile user
1988 J.-B. Ammirati, G. Vargas, S. Rebolledo, R. Abrahami, B. Potin, F. Leyton, and S. Ruiz

seismic event with a local magnitude ML > 2:5 paying par-


ticular attention to the Chilean margin, with rapid characteri-
zation of large magnitude subduction earthquakes for
postseismic response and tsunami alert. During 2017, the
CSN extended the national seismic network with seven
broadband seismometers (including one 100 m deep bore-
hole) permanently deployed at the piedmont of the PC,
around the scarp of the SRF and farther east (Fig. 3 and
Ⓔ Table S1, available in the supplemental content to this
article). The main objective of this permanent instrumenta-
tion is to increase the station concentration in the PC,
improving the detection and characterization of small-
magnitude (M L < 2:5) earthquakes not currently automati-
cally detected by the CSN.
Recent studies explored potential rupture scenarios of the
Figure 3. Map showing the location of the seismic stations used
SRF. Pérez et al. (2014) estimated PGAvalues greater than 0:7g
in this work. The solid line corresponds to the scarp of the SRF as
described in Armijo et al. (2010). Triangles correspond to the CSN based on five random fractal slip distribution models, account-
permanent stations installed prior to 2017. Inverted triangles corre- ing for source directivity effects. Based on transient electro-
spond to stations installed during 2017 as part of the SRF project. magnetic imaging, Estay et al. (2016) evidenced surface
The diamond corresponds to Global Seismographic Network (GSN) segmentation of the SRF and calculated the PGA associated
station Peldehue (PEL). The polygon at the center of the map cor-
with the potential rupture of each segment. Their estimated
responds to the metropolitan area of Santiago. The color version of
this figure is available only in the electronic edition. PGAs are similar to the results of Pérez et al. (2014). The
two aforementioned rupture models consistently show stronger
velocity model calibrated for their study region. Because of accelerations for the hanging wall (HW), east of the SRF scarp.
the sparse station coverage of the Chilean seismic network
at that time, this study pointed out the necessity of having Earthquake Characterization Procedure
a denser array of seismometers deployed in the Cordilleran
sector to improve the detection of smaller events and their The first step of this work consists in analyzing continu-
localizations. Results of Barrientos et al. (2004) reveal a more ous records from the stations shown in Figure 3 and
Ⓔ Table S1 to detect seismic events related to the seismic
clustered seismicity observed at crustal level although not par-
activity of the PC. To make good use of the denser CSN array
ticularly consistent with the geological feature observed at the
around the SRF (Fig. 3), we analyze the continuous wave-
surface, in particular the Pocuro fault zone (a Mesozoic nor-
forms recorded since the deployment of the seven new sta-
mal fault ∼100 km north of Santiago that would have been
tions, between 20 January 2017 and 15 March 2019.
reactivated as a reverse fault during the Neogene) and the
Because the CSN stations record a very large number of
SRF more to the south, in our study area (Fig. 2), two impor-
earthquakes from different sources, a challenge exists in dif-
tant features associated with the uplift of the PC (Charrier
ferentiating crustal events occurring beneath the PC
et al., 2005; Armijo et al., 2010; Farías et al., 2010). (Ⓔ Fig. S3) from the events occurring along the Chilean
Using seismic tomography, Farías et al. (2010) margin or within the subducting slab. For each event, an
improved on the relocalization of the crustal events from the automatic detection and a first location are performed using
CSN catalog between 1980 and 2004. Their findings show a the methodology described in Poiata et al. (2016). This
good correlation between geological observations at the sur- method is based on the statistical characterization of seismic
face and structural reconstruction at depth (Giambiagi, signals recorded by a dense seismic network and the stacking
Ramos, et al., 2003), which suggests that most of the seis- of time-delay functions. The location process of the method
micity is related to mountain building mechanisms respon- allows us to discard all events with hypocenters located out-
sible for the uplift of the Andean Cordillera. Interestingly, the side our study area. For each event detected, we visually
aforementioned seismological studies evidenced clusters of inspect the corresponding traces to remove eventual false
earthquakes beneath the San José (∼33:8° S) and the positives. Because the automatically picked P- and S-wave
Tupungatito (∼33:4° S) volcanoes (Fig. 2). arrival is not always accurate, this step is also a good oppor-
The main task of the CSN consists of the maintenance tunity to manually pick the P- and S-wave first arrivals to
and operation of 105 broadband seismic stations, 128 Global ensure optimized residuals for the final location. When the
Navigation Satellite Systems (GNSS) stations, and 297 traces present a high signal-to-noise ratio, polarity informa-
accelerometers for strong-motion measurement (Barrientos tion for the first P-wave arrival is reported.
and National Seismological Center [CSN] Team, 2018), per- To improve the precision of seismic locations, we need
manently deployed along the country. Among other tasks, an accurate wave velocity model with finer crustal details.
the CSN automatically detects and characterizes every P- and S-wave arrival times of 11,829 earthquakes occurred

Downloaded from https://pubs.geoscienceworld.org/ssa/bssa/article-pdf/109/5/1985/4833931/bssa-2019082.1.pdf


by Univ de Chile user
The Crustal Seismicity of the Western Andean Thrust (Central Chile, 33°–34° S) 1989

in the Chile central region (28°–36° S) and recorded by the square (rms) lower than 1 s, P- and S-wave residuals at each
CSN network over a period of ∼20 yr (Fig. 1) were previously station are averaged. Final maximum-likelihood hypocenters
inverted using a nonlinear approach based on the law of large (Fig. 4 and Ⓔ Table S2) are then estimated taking into account
numbers (Potin, 2016). This step yielded 3D variations of the station delays. We note that in general, the dispersion of time
velocity structure for both P and S waves. Then, we extracted a residuals is low with a standard deviation σ  0:09 and 0.16 s
3D velocity model centered on our region of interest, which for P and S phases, respectively. The highest residuals can be
extends from 33° S to 34.5° S and 69.2° W to 71.6° W, down to observed for stations MT02 (−0:22 s for P phases and
45 km depth and with a grid spacing of 1 km (Ⓔ Fig. S1). −0:31 s for S phases) and MT14 (0.1 and 0.25 s for P and
The events previously detected were located using the S phases, respectively). Station MT14 is located in the urban
NonLinLoc package (Lomax et al., 2000). This nonlinear, area of Santiago (Fig. 3) and presents a relatively high level of
global search method based on the probabilistic formulation anthropogenic noise (Ⓔ Fig. S2), which could be a reason for
of inverse problems described in Tarantola and Valette such high residuals. Station MT02 is one of the quietest sta-
(1982) is particularly well adapted to the use of 3D velocity tions with a remote location in the Coastal Cordillera and in
models because it does not require the calculation of partial general present arrivals with high signal-to-noise ratios (Fig. 3
derivatives (otherwise very difficult to perform using linear Ⓔ Fig. S2). High residuals observed for this station could be
approaches). The locations obtained in this work are repre- related to some inadequacy of our 3D velocity model in the
sented by a probability density function (PDF) that produces vicinity of this particular station. The rms associated with the
more comprehensive uncertainty estimations. The minimum final maximum-likelihood locations is 0.1 s in aver-
requirements for an event to be located were set to a mini- age (σ  0:07 s).
mum of eight picks (both P and S) and at least two S-wave Location errors are estimated from the 68% confidence
picks. Once located, we estimated the local magnitude (M L ) ellipsoid as computed from the samples of the location PDF.
based on shear-wave maximum amplitude. The average horizontal error is 2.3 km (σ  1:5 km) and the
Finally, focal mechanisms and corresponding compres- vertical error is about 3.0 km (σ  2 km). In general, we
sion (P) and tension (T) axes were determined from P-wave obtained hypocenters with smaller (below average) vertical
first-motion polarities using the grid-search algorithm FPFIT error for events with locations closer to a seismic station.
(Reasenberg and Oppenheimer, 1985). We used maximum- Increasing azimuthal gap seems to be associated with
likelihood hypocenters and corresponding ray takeoff angles increasing horizontal error (Fig. 5).
from the 3D locations as an input. Because the seismicity Twenty-nine events from this work were also detected
detected is very low in magnitude (in general lower than and located by the CSN (Fig. 5d). For the central Chile
ML < 2:5), the signal-to-noise ratio is often poor, in particu- region, the CSN uses an improved 1D, three layers over a
lar for stations located close to the urbanized area (Fig. 3 and half-space velocity model to locate (M L > 2–2:5) earth-
Ⓔ Fig. S2), which makes the P-wave polarity hard to read quakes automatically (Barrientos et al., 2004; Barrientos and
for these stations. Thus, we determined fault plane solutions National Seismological Center [CSN] Team, 2018). The rms
for events with a minimum of six good first-motion polarity for the CSN events is no larger than 0.3 s and the location
readings (N R ; Ⓔ Fig. S4). errors are in general low (comparable to the ones obtained in
Under the main assumptions of a uniform regional stress this work only with a slightly higher standard deviation).
tensor in the crust, within our study region and that the slip is However, if we compare the CSN hypocenter solutions with
parallel to the direction of the tangential traction (Wallace, the ones obtained in this work, we notice considerable dis-
1951; Bott, 1959), it is possible to find the three principal crepancies, especially in the vertical direction. The hypocen-
stress directions that will most closely match the focal ters obtained using the 3D model seem to be shallower and
mechanism observations by performing formal stress inver- more gathered around 10 km depth in comparison with the
sion (FSI; Hardebeck and Michael, 2006; Martinez-Garzón CSN. The difference tends to increase as we look to the east.
et al., 2014). Principal stress axis uncertainties are evaluated Beneath the FC (∼69:6° W), the difference in hypocenter
by randomly resampling the focal mechanisms data (boot- depths can reach 15 km (Fig. 5e). The CSN 1D layered
straping). In this work, FSI is performed for low quality model has been calibrated for the Andes of central Chile
(N R > 6), intermediate quality (N R > 8), and higher quality (Barrientos et al., 2004). We note that the difference in focal
(N R > 10) focal mechanisms. Only seismic locations with a depth is higher for events located in the back-arc region of
maximum azimuthal gap of 220° have been considered. Argentina, where the real velocity variations might be not
well represented by the velocity model. We believe that our
3D velocity model allows more realistic raytracing, hence
Results more accurate hypocenter locations.
Looking at the distribution of hypocenters (Fig. 4), we
Earthquake Locations
observe that the seismicity is mainly located along two
We obtained 917 probabilistic locations corresponding to north–south stripes. The first stripe extends beneath the PC
the crustal seismicity recorded mainly beneath the PC between (70°–70.2° W) between latitudes 33.2° S and 34.1° S. The
January 2017 and March 2019. For events with a root mean depth associated with this seismicity appears quite constant

Downloaded from https://pubs.geoscienceworld.org/ssa/bssa/article-pdf/109/5/1985/4833931/bssa-2019082.1.pdf


by Univ de Chile user
1990 J.-B. Ammirati, G. Vargas, S. Rebolledo, R. Abrahami, B. Potin, F. Leyton, and S. Ruiz

Figure 4. (Top) Epicentral distribution of the seismic events characterized in this work from 27 January 2017 to 15 March 2019. The
polygon shows the limits of the metropolitan area of Santiago. Squares mark the location of the seismic stations used for the location (see
Fig. 3 for all station codes). (a–f) Solid lines mark the emplacements of the cross sections. Hypocenters located 0.1° apart from the lines are
projected. (Bottom) Corresponding cross sections showing (a–f) the distribution at depth of the seismicity reported in this work. Circled
numbers identify clusters of seismicity not directly related to the tectonics of the west Andean thrust: (1) Santiago cluster, (2) Los Bronces
open pit mine, and (3) Tupungatito volcano. The color version of this figure is available only in the electronic edition.

at ∼10 km (5–15 km). The second stripe can be observed more 20–30 km (Fig. 4). This seismicity has been evidenced in
to the west consistently aligned with the western flank of the PC previous studies although the mechanisms behind it remain
(∼70:4° W). Hypocenters in this sector are 10–20 km deep. quite hypothetical (Leyton et al., 2009). During August
Interestingly, we observe some clustered seismicity 2017, the CSN network recorded a seismic swarm (Fig. 4)
beneath the Central Depression (∼33:6° S) at depths of about located beneath the Tupungatito volcano (Fig. 2) at

Downloaded from https://pubs.geoscienceworld.org/ssa/bssa/article-pdf/109/5/1985/4833931/bssa-2019082.1.pdf


by Univ de Chile user
The Crustal Seismicity of the Western Andean Thrust (Central Chile, 33°–34° S) 1991

Figure 5. (a) Variability of the vertical error depending on earthquake location. Note how lower vertical error decreases when hypo-
centers are closer to a seismic station. (b) Similar to (a). The map is showing the variation of the horizontal error. This time, higher errors seem
to be associated with increasing azimuthal gap (c). (d) Epicenter distribution of some of 29 earthquakes located in this work and by the CSN.
(e) Depth distribution for the 29 earthquakes located in this work and by the CSN. The solid line corresponds to the Chile–Argentina
international border. The color version of this figure is available only in the electronic edition.

10–15 km depth. Furthermore, we can observe a small clus- can occur on preexisting zones of weakness in directions not
ter of seismicity at shallow depths (0–5 km) beneath the Los always geometrically consistent with regional stress directions
Bronces open pit mine. These clusters are not related to the (McKenzie, 1969). Because the seismicity found in this work
tectonics of the western Andean thrust; hence, we do not con- is low in magnitude, we expect the corresponding focal mech-
sider them in our further interpretations. anisms to be associated with small fractures with a broad
Sections across the study area (Fig. 4) show that the range of orientations rather than larger regional structures
seismicity depth presents a general westward increase. globally aligned with observed geological features.
Seismicity in the Argentine back-arc is limited to the north- Considering this, the focal mechanism parameters (strike,
east of our study area and can be observed at shallow depth dip, and rake) are not limited by regional structural observa-
(although we acknowledge that the location errors are higher tions as seen in previous works (e.g., Pérez et al., 2014).
in this sector due to a higher azimuthal gap). We obtained a total 104 focal mechanism solutions from
The local magnitudes estimated in this work are in gen- P-wave first-motion polarities (Fig. 6 and Ⓔ Table S3). Fifty-
eral lower than M L < 2:5 (Ⓔ Table S2). However, In the two of them were determined using a minimum of N R > 8
Argentine back-arc, the magnitude can be higher with events polarity readings and 27 has a minimum of N R > 10 polarity
with magnitude M L > 3:0. readings. The limited number of polarity data and their weak
coverage of the focal sphere did not permit to accurately con-
strain the fault plane orientations for each individual event.
Focal Mechanism Solutions
Thus, we do not relate them with geological structures but
A single focal mechanism only constrains the stress direc- rather look at the set of solutions as a whole.
tion within the dilatational quadrant, which implies that vari- In general, our solutions (Fig. 6a and Ⓔ Table S3) cover
ous focal mechanisms are necessary to further constrain the a large variety of focal mechanisms. However, we observe a
regional stress directions. In a rock volume under stress, slip major concentration of T axis closer to the center of the focal

Downloaded from https://pubs.geoscienceworld.org/ssa/bssa/article-pdf/109/5/1985/4833931/bssa-2019082.1.pdf


by Univ de Chile user
1992 J.-B. Ammirati, G. Vargas, S. Rebolledo, R. Abrahami, B. Potin, F. Leyton, and S. Ruiz

right-lateral strike-slip component compatible with the focal


mechanism of intermediate magnitude earthquake located at
shallow depths, in the PC sector (Fig. 1).

Implications for Regional Tectonics


Both improved hypocentral locations and regional stress
tensor obtained in this work can be discussed in the frame-
work of the regional tectonics in our study area.
The hypocenter distribution (Fig. 4) seems to coincide
with tectonic and geological observations at the surface
(Fig. 2). We observe that an important amount of the shallow
seismicity lies beneath the highly deformed Mesozoic
sequences of the PC also known as the Aconcagua fold and
thrust belt (AFTB). Around 33.5° S, this ∼20 km wide fea-
ture has been described as a thin-skinned thrust belt with
east-vergent structures, overthrusting Neogene intramoun-
tain basins. The timing of the deformation suggests that this
feature contributed to the accommodation of the Andean
orogeny between ∼20 and ∼10 Ma (Giambiagi, Alvarez,
et al., 2003). The basal décollement of the AFTB has been
estimated by Giambiagi, Alvarez, et al. (2003) to be 5–10 km
deep, which is quite consistent with the distribution of our
hypocenters for this area (Fig. 7). However, recent re-evalu-
ation of the structural configuration by Riesner et al. (2018)
suggests that the AFTB roots onto a very shallow, 2–3 km
deep décollement level, well above the seismicity level high-
lighted in the present work.
Figure 6. (a) Inferior hemisphere projection of the P (compres-
sion) and T (tensional) axes for the focal mechanisms found in this
Two controversial models have been proposed to
study with a minimum of 6, 8, and 10 polarity readings, respec- explain the tectonic evolution of the PC and the FC more to
tively. (b) Formal stress inversion (FSI) results. The crosses corre- the east. Giambiagi et al. (2015) proposed that the AFTB
spond to the best regional stress orientations. The data have been accommodated crustal shortening in the PC during the
randomly resampled (2000 times) for uncertainty estimations. The Miocene and then migrated to the east, affecting the
dots show the distribution of the bootstrap samples. The color
version of this figure is available only in the electronic edition.
Permian–Triassic sequences of the FC. This has been con-
strued as two successive east-vergent thrusts with two sep-
sphere and P axis more distributed closer to the edges of the arate décollement levels at 5–10 and 15–20 km depths,
respectively. The total estimated shortening associated with
focal sphere suggesting compressive regional stress.
these structures is ∼47 km (Giambiagi and Ramos, 2002).
FSI results (Fig. 6b) show a general northeast–southwest
Alternatively, Armijo et al. (2010) consider the AFTB as
orientation of the principal stress axis (σ 1 ). The minimum
a secondary feature that accommodated a maximum of
stress axis (σ 3 ) appears well constrained at the center of
10 km of crustal shortening and have then been carried over
the focal sphere, which suggests a northeast–southwest-ori-
the basement of the FC by a deep west-vergent ramp
ented compressional crustal stress state for our study region.
throughout a continuous crustal shortening accommodation
These results are consistent with the regional velocity field during the Neogene. The important seismicity observed in
inferred from GNSS data (Métois et al., 2016). the present work (Figs. 4 and 7) beneath the AFTB is com-
In detail, the uncertainty associated with principal stress patible with the uplift of the FC basement as mentioned by
orientations does not appear to be sensitive to data quality. Riesner et al. (2018) although they imply that this structure
The bootstrap samples are tightly surrounding the best sol- was only active during the late Middle Miocene (∼11 Ma).
ution especially when the number of polarity readings is low Our results thus suggest that quite an important fraction
(6 < N R < 8). We believe that the larger dispersion of the (∼70%) of the seismicity in the west Andean thrust occurs
bootstrap samples for good quality events is due to the at the transition between the PC and the FC basements,
smaller size of the corresponding dataset (only 27 focal around ∼10 km depth (Fig. 7).
mechanisms were inverted in this case). Another interesting As previously mentioned, hypocenters consistently align
observation is that the stress tensor seems to rotate around the with the western flank of the PC (Figs. 2 and 4), right beneath
principal stress axis (σ 1 ) as we perform FSI using fewer focal the Farellones plateau (70.3°–70.5° W) at a depth of ∼15 km.
mechanisms. This rotation is introducing some north–south In this sector, the Pre-Andean basement and the overlying

Downloaded from https://pubs.geoscienceworld.org/ssa/bssa/article-pdf/109/5/1985/4833931/bssa-2019082.1.pdf


by Univ de Chile user
The Crustal Seismicity of the Western Andean Thrust (Central Chile, 33°–34° S) 1993

Figure 7. Integrated cross section showing the depth distribution and the normalized density of the seismicity characterized in this study
between 33° S and 34.5° S. Solid lines are structures inferred from geological observations (Riesner et al., 2018). Dashed lines correspond to
major structures inferred from our seismic results. Dotted lines delimit the seismogenic zone associated to the tectonics of the study region.
The circled numbers refer to clusters of seismicity not directly related to the tectonics west Andean thrust (see Fig. 4). The color version of
this figure is available only in the electronic edition.

Abanico and Farellones formations as well as Quaternary of at least 50 km. This length value is more consistent with
intrusive bodies seem to be rather deformed and uplifted by empirical scaling laws between 4.7 and 4.9 slip estimated by
deep-cored structures sometimes referred as west Andean fold Vargas et al. (2014) and the rupture length at the surface
and thrust belt (WAFTB); Riesner et al., 2017). The PC is (Wells and Coppersmith, 1994). However, we acknowledge
bounded to the west by east-dipping faults that mounted the that the amount slip measured at the surface is not always rep-
aforementioned Cenozoic formations above the Quaternary resentative of the magnitude, because the rupture is not nec-
sediments of the Central Depression resulting in a very abrupt essarily uniform along the fault plane. As an example, the
mountain front (Fig. 2). Farther east, a series of anticlines and 1999 Mw 7.6 Chi-Chi earthquake (Shin and Teng, 2001) was
synclines was developed as the underlying structures (former characterized by a slip and surface rupture length way above
WAFTB) accommodated the Andean compression for the past the values expected from magnitude scaling relationships.
20–25 Ma. According to the west-verging model (Armijo In the Andes of central Chile, most of the tectonic fea-
et al., 2010; Riesner et al., 2017), the WAFTB would be the tures observed at the surface present a north–south general
main structure responsible for the building of the Andes orientation. This is particularly true for the three main afore-
between 33° and 35° S. Recent quantification of the WAFTB mentioned structures: the AFTB in the eastern PC, the
kinematics from structural and geochronological observations WAFTB in the western PC, and the SRF (Fig. 2). These
(Riesner et al., 2017) allowed to estimate the depth of the structures accommodated the Andean shortening in a very
décollement level associated with the WAFTB at 12–15 km. clear east–west direction. However, the regional stress tensor
The hypocenter distribution found along the western front of displays a northeast–southwest orientation, parallel to the
the PC, lying at ∼15 km depth is compatible with this struc- convergence orientation of the Nazca plate. This observation
ture (Fig. 4). Nearby Santiago de Chile, the west Andean front raises the question of how the upper plate accommodates
is characterized by the presence of the SRF. The seismicity northeast–southwest motion induced by the subduction
associated with the WAFTB appears quite parallel to this dynamics with north–south structure? Geological observa-
structure (Fig. 4) although we do not observe any hypocenters tions mainly support the idea of reverse faulting. As an exam-
on the SRF fault plane, at shallow levels, between the surface ple, the SRF on the PC piedmont previously interpreted as a
and the base of the ramp. Paleoseismic studies suggest that at normal fault (Brüggen, 1950) has been recently studied more
least two ruptures occurred along the SRF during the past in details driving to the conclusion that it rather corresponds
17,000 yr (Vargas et al., 2014) and were associated to 7:2 > to the expression of a west-vergent reverse structure thrusting
Mw > 7:5 earthquakes. We believe that such high-magnitude the Miociene volcanic rocks of the Abanico Formation over
events would occur when the entire ramp is ruptured. the Quaternary sediments of the Central Depression (Armijo
Although the SRF scarp at the surface was mapped as a et al., 2010; Vargas et al., 2014). More to the east, the west-
∼30 km long structure (Armijo et al., 2010), the seismicity verging folds that characterize the WAFTB present clear
(found in this work) at its base clearly extends farther south. north–south axial planes and suggest east–west deformation
Considering this observation, there is a possibility that the (Riesner et al., 2017). The same observation would apply to
SRF plane is extending southward as well, totaling a length the AFTB (Armijo et al., 2010; Riesner et al., 2018).

Downloaded from https://pubs.geoscienceworld.org/ssa/bssa/article-pdf/109/5/1985/4833931/bssa-2019082.1.pdf


by Univ de Chile user
1994 J.-B. Ammirati, G. Vargas, S. Rebolledo, R. Abrahami, B. Potin, F. Leyton, and S. Ruiz

The Las Melosas earthquake is the largest crustal event They obtained PGA values between 0.7 and 0:8g around the
instrumentally recorded in the central Andes. It occurred on 4 scarp of the SRF.
September 1958 in the PC, ∼60 km southeast from Santiago. Estay et al. (2016) focused their work on improving the
The hypocenter was located at around 10 km depth, beneath near-surface fault geometry. They found evidence of at least
the locality of Las Melosas where the earthquake was felt four different subsegments of ∼10 km length, along the SRF.
with the strongest intensity (IX on the Mercalli scale) accord- They concluded that those segments would be activated inde-
ing to Sepúlveda et al. (2008). Using teleseismic waveform pendently resulting in four different rupture scenarios that
modeling and moment tensor inversion, Alvarado et al. involve earthquakes of magnitudes 6:2 < Mw < 6:7.
(2009) were able to further constrain the corresponding Corresponding accelerations were computed for a grid of
source parameters. In particular, they calculated a moment 1 km spacing. V S30 and basin depth variations were taken into
magnitude of Mw 6.3 associated with a strike-slip (right-lat- account. Despite the even more conservative magnitude range,
eral) focal mechanism. The optimal source parameters were PGA estimations are consistent with the results of Pérez
obtained for a depth of 8 km, which is consistent with pre- et al. (2014).
vious studies based on local seismograms analysis (Pardo As mentioned in the previous section, based on the seis-
and Acevedo, 1984). To explain such a mechanism, micity obtained in the present work, in particular the seismic-
Alvarado et al. (2009) proposed that the Las Melosas event ity associated with the western flank of the PC (Figs. 4 and 7),
could have been associated with the activation of east–west we consider the possibility that the SRF extends farther south,
structures accommodating differences in shortening rate down to ∼33:7° S totaling a surface length of 50 km. The ramp
north from 33° S although no surface rupture were observed of the SRF would connect with the western PC seismicity at
to corroborate this hypothesis. Another possible explanation ∼10 km depth. Thus, we estimate an east-dipping fault plane
is that the strike-slip component induced by the north–east (∼34°) with an area of S  50 × 20km2 (Fig. 8). Seismic
regional stress is accommodated by the existent north–south moment (M0 ) and moment magnitude (M w ) can be estimated
crustal structures (Fig. 2) although in this case, the focal by relating the shear modulus (μ), the rupture area (A), and the
mechanism of the Las Melosas earthquake does not match amount of slip produced by the earthquake (d) (Kanamori,
the regional stress tensor (Fig. 6). Two other crustal events 1977). Our scenario considers a seismic event that would
occurred in 1987 and 2001 (Alvarado et al., 2005) show rupture the entire fault plane. Hence, taking a shear modulus
focal mechanisms compatible with this idea (Fig. 1). μ  40 GPa (Ji et al., 2010), a surface area S  1 × 106 m2
Interestingly, these strike-slip events seem to be located in and a slip (assumed uniform along the rupture) of d  4:9 m
the eastern PC and FC, indicating that the regional north– (Vargas et al., 2014) we obtain a seismic moment of
east-oriented principal stress could be accommodated in this M0  1:96 × 1020 N · m (M w 7.46).
sector. More intermediate magnitude (M > 5) events in this We use ground-motion prediction equations derived from
sector would be necessary to validate this particular point. the Next Generation Attenuation (NGA) database (Pacific
Earthquake Engineering Research Center [PEER], 2019) to
compute PGAs in our study area (Abrahamson et al., 2014;
Implications for Seismic Hazard Graizer and Kalkan, 2016; hereafter, ASK14 and GK15,
respectively). The NGA database includes ground-motion
SRF Rupture Scenario
observations for a large number of earthquakes, which allows
The scarp of the SRF has been identified bordering the the calculation of PGA median value and standard deviation.
eastern side of the metropolitan area of Santiago. It extends For this reason, we assume that PGA variations due to source
from the Cerro Alvarado, on the north bank of the Mapocho effects are taken into account through the dispersion of the
river (−33:35° S) to the south of the Maipo river at ∼33:6° S empirical data. The HW effect has been defined as a system-
(Armijo et al., 2010). Previous SRF rupture scenarios (Pérez atic amplification of ground-motion amplitudes measured on
et al., 2014; Estay et al., 2016) are based on approximation the HW (Abrahamson and Somerville, 1996). Unlike GK15,
of Armijo et al. (2010) of fault plane geometry. the ASK14 model allows to specifically quantify this effect.
Pérez et al. (2014) estimated the PGA from five random For this reason, we compute PGA variations using both mod-
composite fractal slip distribution scenarios and calculated els to explore their similarities and differences.
the corresponding PGA within a 122-node grid. This Our SRF scenario is based on the most recently available
approach accounts for the directivity effect associated with geological and seismological knowledge. It considers the
the rupture propagation. This effect strongly influences the rupture of every point on the fault plane (Fig. 8), which does
wave amplitudes depending on the angle between the not mean that the rupture would be uniform along the entire
receiver and the rupture propagation main direction. On the fault plane but that we consider every possible source–
other hand, they neglected site parameters that could locally receiver distances. We compute PGAs for a grid of ∼900 m
influence the ground acceleration such as soil qualifications (30 arcsec) spacing for which each node is characterized with
(V S30 ) and/or basin thickness variations. Their rupture sce- elevation (h), V S30 and the closest distance to the rupture
narios were computed for a conservative magnitude M w 6.9 (rmin ). The distance to the rupture is calculated considering
earthquake rupturing an S  30 × 16 km2 SRF fault plane. the topography and the depth of the fault plane. Thus, for

Downloaded from https://pubs.geoscienceworld.org/ssa/bssa/article-pdf/109/5/1985/4833931/bssa-2019082.1.pdf


by Univ de Chile user
The Crustal Seismicity of the Western Andean Thrust (Central Chile, 33°–34° S) 1995

(footwall) is filled with relatively low V S30 (450–500 ms−1 )


detrital alluvium in comparison to the HW, mostly composed
by higher V S30 (>1200 ms−1 ) rocks from the Abanico for-
mation (Fig. 2 and Ⓔ Fig. S5). This difference in V S30 seems
to compensate the HW effect. However, our highest PGA
values are observed on the HW north of our study area
(∼33:4°S) and more to the south (∼33:6°S) where the
Mapocho and Maipo rivers (Figs. 2 and 9) open to the
Santiago basin. These sectors are characterized by low
V S30 (∼500 ms−1 ) and are located close to the SRF scarp,
Figure 8. SRF plane geometry taken into account for a major hence the strong PGAs observed locally. To the northwest
rupture scenario. The solid line corresponds to the scarp of the SRF of the Santiago metropolitan area (Fig. 9), the presence of
as described in Armijo et al. (2010). The dashed line is the southern volcanic ashes (Pudahuel ignimbrites) is characterized by
extension of the SRF inferred from the crustal seismicity reported in a very low V S30 (∼150 ms−1 ) and contributes to a locally
this work (see the Results section). The rupture of the entire SRF higher PGA (0:3  0:15g) in this sector despite being
fault plane would generate an earthquake of magnitude Mw 7.5. The
metropolitan area of Santiago (white polygon) is home to more than located further from the fault scarp. This observation is par-
∼6:5 million inhabitants. The color version of this figure is available ticularly visible on the GK15 model (Fig. 9a and Ⓔ Fig. S5).
only in the electronic edition. Overall, our models show that ground-motion attenua-
tion with distance is lower than observed in Pérez et al.
the foot wall (west of the SRF scarp) rmin corresponds to the (2014) and Estay et al. (2016). On the footwall, 10 km from
closest distance to the scarp whereas some points on the HW the SRF scarp, our models show PGA average values of
can be closer to the underlying fault plane. The shear-wave 0:4–0:5g, in which the work mentioned earlier obtained val-
velocity to a depth of 30 m (V S30 ) has been estimated in the ues of ∼0:25 and ∼0:3g, respectively. This difference in
study area by Leyton et al. (2011) based on soil and rock ground-motion attenuation can be related to the higher mag-
characteristics from local geology. We considered basin nitude earthquake modeled in this study.
thicknesses (b) from Yáñez et al. (2015), which varies Values and distribution of PGAs estimated in this study
between 0.2 km to the south and 0.5 km to the northeast of are consistent with ground-motion measurements correspond-
the Santiago basin. The Joyner–Boore distance (closest dis- ing to the 1999 M w 7.6 Chi-Chi earthquake, a case similar to
tance to the surface projection of the rupture plane: RJB ) is our SRF scenario in terms of tectonic setting although with a
also calculated, as it is required by ASK14. The anelastic slightly higher magnitude and a rupture length of nearly 90 km
attenuation factor (Q0 ) is required to constrain far source (Shin and Teng, 2001). Chi-Chi generated PGA values greater
attenuation in the GK15 model (which in our case will only than 0:9g close to the rupture scarp, on the HW and PGA val-
have little effect). We set Q0  450 based on Ji et al. (2010). ues of 0:3g at ∼25 km to the west, on the footwall. Another
comparable example would be the 2011 M w 6.2 Christchurch
earthquake with recorded PGA greater than 0:8g on the HW
Results and Discussion
although with a rapid decrease at short distances
PGAs computed in this work are shown in Figure 9. In (PGA  0:25g recorded at 10 km from the epicenter), prob-
general, both models show quite similar results close to the ably due to the lower magnitude compared to our study. Also,
SRF scarp with maximum PGA of 0:8  0:4g and the Christchurch area is characterized by poorer quality soils
0:7  0:3g for the ASK14 and GK15 models, respectively. than observed in the Santiago basin. Unlike, the SRF or the
We observe that footwall PGAs obtained from ASK14 1999 Chi-Chi case, the fault responsible for the 2011
decrease more rapidly with distance compared to the GK15 Christchurch earthquake was blind, which means that no sur-
model. Values obtained in this work are consistent with Pérez face rupture was observed.
et al. (2014) and Estay et al. (2016) although our rupture Shallow crustal earthquakes this size occurring on reverse
scenario is considering a much bigger earthquake (Mw 7.5). faults are expected to generate surface rupture, which was the
However, our results show quite a different repartition of case for the Chi-Chi earthquake and locally resulted in a ∼9 m
PGAs compared to the aforementioned works. Close to the scarp. This amount of slip close to the surface reminds us that
SRF scarp, the first-order parameter that controls the amount the slip distribution can be quite heterogeneous along the rup-
of expected acceleration appears to be the V S30 , a parameter ture and that patches of high slip can locally increase PGA
closely related to the observed geology in the study area values. Considering the 4.9 m surface displacement observed
(Leyton et al., 2011). Pérez et al. (2014) and Estay et al. by Vargas et al. (2014), there is no doubt that the rupture of the
(2016) estimated stronger PGAs systematically located on SRF will reach the surface. In this case, differential ground
the HW. Although we specifically considered the amplifica- movements would be particularly detrimental for any build-
tion caused by the HW effect, our results (Fig. 9) show ings or infrastructure constructed above and right next to the
relatively higher PGAs mostly characterizing the footwall. rupture area. We do not discard the occurrence of smaller
The reason for this observation is that the Santiago basin events, large enough to cause damage but with a slip

Downloaded from https://pubs.geoscienceworld.org/ssa/bssa/article-pdf/109/5/1985/4833931/bssa-2019082.1.pdf


by Univ de Chile user
1996 J.-B. Ammirati, G. Vargas, S. Rebolledo, R. Abrahami, B. Potin, F. Leyton, and S. Ruiz

Figure 9. Peak ground acceleration (PGA) estimations corresponding to our rupture scenario for the SRF calculated from the (a) GK15
and (b) ASK14 models, following Graizer and Kalkan (2016) and Abrahamson et al. (2014), respectively. The solid white line corresponds to
the SRF scarp considered in this case. Both models clearly show high-PGA values on the footwall, east from the scarp of the SRF. The
strongest accelerations are observed on the hanging wall (HW) in areas filled with sedimentary material carried by the Mapocho river to the
north and then by the Maipo river to the south (Fig. 2). This observation is particularly clear on the (b) ASK14 model that specifically
accounts for the HW effect (see the Implications for Seismic Hazard section). Variations of PGA mean values for the GK15 and
ASK14 models along cross section AB are shown within 1 standard deviation in (c) and (d), respectively. The dashed line shows topography
variations along cross section AB. The color version of this figure is available only in the electronic edition.

Downloaded from https://pubs.geoscienceworld.org/ssa/bssa/article-pdf/109/5/1985/4833931/bssa-2019082.1.pdf


by Univ de Chile user
The Crustal Seismicity of the Western Andean Thrust (Central Chile, 33°–34° S) 1997

distribution that would not rupture up to the surface. Finally, N° 41, 20 June 2016 (Monitoreo sismico y potencial sismogénico de la Falla
strong accelerations related to the SRF rupture could trigger San Ramón). The authors thank Sergio Barrientos and the Chilean
Seismological Center (CSN) team (in particular Sebastián Arriola) for the
landslides and rock falls in areas characterized by high easy access to the continuous waveforms and station metadata used in the
topography. present work. J.-B. A. is thankful to Chelsea Mackaman-Lofland for fruitful
discussions about the tectonics of the central Andes. J.-B. A., G. V., and S. R.
Conclusions are thankful to Ruben Boroschek for his feedback on peak ground acceler-
ation (PGA) calculations. The authors are grateful to Associate Editor Mark
In this study, we analyzed 26 months (27 January 2017 to W. Stirling, as well as Gregory De Pascale and another anonymous reviewer
for their very constructive comments on the original article.
15 March 2019) of continuous waveforms recorded by 17
broadband seismometers and obtained 917 probabilistic loca-
tions of seismic events associated to the west Andean thrust References
between 33° and 34° S. Our results evidence seismic activity
mainly distributed beneath the eastern PC between 5 and Abrahamson, N. A., and P. G. Somerville (1996). Effects of the hanging wall
15 km depths and beneath the western PC between 10 and and footwall on ground motions recorded during the Northridge earth-
quake, Bull. Seismol. Soc. Am. 86, no. 1B, S93–S99.
20 km depths. The former imply that crustal shortening affect- Abrahamson, N. A., W. J. Silva, and R. Kamai (2014). Summary of the
ing the central Andes would be concentrated at the transition at ASK14 ground motion relation for active crustal regions, Earthq.
depth between the PC and the FC at depth. The latter is con- Spectra 30, no. 3, 1025–1055, doi: 10.1193/070913EQS198M.
sistent with the west Andean thrust and would be associated Allmendinger, R. W., D. Figueroa, D. Snyder, J. Beer, C. Mpodozis, and B.
with the SRF showing that the west Andean front is active. L. Isacks (1990). Foreland shortening and crustal balancing in the
Andes at 30°S latitude, Tectonics 9, no. 4, 789–809, doi: 10.1029/
Focal mechanisms obtained in this work allowed to constrain TC009i004p00789.
a northeast–southwest direction of the principal stress axis, in Alvarado, P., S. Barrientos, M. Saez, M. Astroza, and S. Beck (2009). Source
good agreement with GNSS velocity measurements. This dis- study and tectonic implications of the historic 1958 Las Melosas crustal
cordance between the principal stress direction and the north– earthquake, Chile, compared to earthquake damage, Phys. Earth Planet.
south orientation of the structure would explain strike-slip In. 175, nos. 1/2, 26–36, doi: 10.1016/j.pepi.2008.03.015.
Alvarado, P., S. Beck, G. Zandt, M. Araujo, and E. Triep (2005). Crustal
mechanisms associated with shallow, intermediate magnitude deformation in the south-central Andes back-arc terranes as viewed
earthquakes in the PC and FC. The crustal seismicity charac- from regional broadband seismic waveform modeling, Geophys. J.
terized in this work has been used to better constrain the geom- Int. 163, no. 2, 580–598, doi: 10.1111/j.1365-246X.2005.02759.x.
etry of the SRF and propose an improved rupture scenario Armijo, R., R. Rauld, R. Thiele, G. Vargas, J. Campos, R. Lacassin, and E.
involving a characteristic Mw 7.5 earthquake. Expected Kausel (2010). The west Andean thrust, the San Ramón fault, and the
seismic hazard for Santiago, Chile, Tectonics 29, TC2007, doi:
PGA in Santiago could be stronger than 0:8g close to the scarp 10.1029/2008TC002427.
of the SRF and greater than 0:3g for most of the metropolitan Barrientos, S., and National Seismological Center (CSN) Team (2018). The
area. Considering this scenario, surface rupture would be very seismic network of Chile, Seismol. Res. Lett. 89, no. 2A, 467–474, doi:
likely, hence the necessity to severely restrict (or prohibit) any 10.1785/0220160195.
kind of new construction in the proximity of the SRF scarp. Barrientos, S., E. Vera, P. Alvarado, and T. Monfret (2004). Crustal seismic-
ity in central Chile, J. South Am. Earth Sci. 16, 759–768, doi: 10.1016/
j.jsames.2003.12.001.
Data and Resources Beck, S., S. Barrientes, E. Kausel, and M. Reyes (1998). Source character-
istics of historic earthquakes along the central Chile subduction zone,
The continuous waveforms used in this study can be J. South Am. Earth Sci. 11, 115–129.
downloaded from the Incorporated Research Institutions for Beyreuther, M., R. Barsch, L. Krischer, T. Megies, Y. Behr, and J.
Seismology (IRIS) platform at https://www.iris.edu/hq/. Data Wassermann (2010). ObsPy: A Python toolbox for seismology,
Seismol. Res. Lett. 81, no. 3, 530–533, doi: 10.1785/gssrl.81.3.530.
preparation and processing were performed using the ObsPy,
Boroschek, R. L., V. Contreras, D. Y. Kwak, and J. P. Stewart (2012). Strong
Pyrocko, and Seismic Analysis Code (SAC) packages ground motion attributes of the 2010 Mw 8.8 Maule, Chile, earth-
(Goldstein et al., 2003; Beyreuther et al., 2010; https:// quake, Earthq. Spectra 28, no. S1, S19–S38, doi: 10.1193/1.4000045.
pyrocko.org). Some of the figures shown in this article were Bott, M. H. P. (1959). The mechanics of oblique slip faulting, Geol. Mag. 96,
made using the Generic Mapping Tools (GMT) package 109–117.
Brüggen, J. (1950). Fundamentos de la geología de Chile, Instituto
(http://www.soest.hawaii.edu/gmt/; Wessel and Smith, 2006).
Geográfico military (IGM), Santiago, Chile (in Spanish).
Global Seismographic Network (GSN) station information Charrier, R., M. Bustamante, D. Comte, S. Elgueta, J. J. Flynn, N. Iturra, N.
(Fig. 3) can be found at https://earthquake.usgs.gov/ Muñoz, M. Pardo, R. Thiele, and A. R. Wyss (2005). The Abanico
monitoring/gsn/. Figure 1 shows three crustal events from the extensional basin: Regional extension, chronology of tectonic inver-
Global Centroid Moment Tensor (CMT) catalog (https:// sion and relation to shallow seismic activity and Andean uplift,
Neues Jahrb, Geol. Palaeontol. Abh. 236, 43–77.
www.globalcmt.org/CMTsearch.html). All websites were last
Cifuentes, I. L., and P. G. Silver (1989). Low frequency source character-
accessed on July 2019. istics of the great 1960 Chilean earthquake, J. Geophys. Res. 94, 643–
663, doi: 10.1029/JB094iB01p00643.
Acknowledgments Cristallini, E. O., and V. A. Ramos (2000). Thick-skinned and thin-skinned
thrusting in the La Ramada fold and thrust belt: Crustal evolution of the
This work was funded by the Chilean Oficina Nacional de Emergencia high Andes of San Juan, Argentina (32°SL), Tectonophysics 317, 205–
del Ministerio del Interior y Seguridad Pública (ONEMI) under resolution 235, doi: 10.1016/S0040-1951(99)00276-0.

Downloaded from https://pubs.geoscienceworld.org/ssa/bssa/article-pdf/109/5/1985/4833931/bssa-2019082.1.pdf


by Univ de Chile user
1998 J.-B. Ammirati, G. Vargas, S. Rebolledo, R. Abrahami, B. Potin, F. Leyton, and S. Ruiz

Ekström, G., M. Nettles, and A. M. Dziewonski (2012). The global CMT Martinez-Garzón, P., G. Kwiatek, M. Ickrath, and M. Bohnhoff (2014).
project 2004-2010: Centroid-moment tensors for 13, 017 earthquakes, MSATSI: A MATLAB package for stress inversion combining solid
Phys. Earth Planet. In. 200/201, 1–9, doi: 10.1016/j.pepi.2012.04.002. classic methodology, a new simplified user-handling and a visualiza-
Estay, N. P., G. Yáñez, S. Carretier, E. Lira, and J. Maringue (2016). Seismic tion tool, Seismol. Res. Lett. 85, no. 4, 896–904, doi: 10.1785/
hazard in low slip rate crustal faults, estimating the characteristic event 0220130189.
and the most hazardous zone: Study case San Ramón fault, in Southern McKenzie, D. P. (1969). The relation between fault plane solutions for earth-
Andes, Nat. Hazards Earth Syst. Sci. 16, 2511–2528, doi: 10.5194/ quakes and the directions of the principal stresses, Bull. Seismol. Soc.
nhess-16-2511-2016. Am. 59, no. 2, 591–601.
Farías, M., D. Comte, R. Charrier, J. Martinod, C. David, A. Tassara, F. Métois, M., C. Vigny, and A. Socquet (2016). Interseismic coupling, mega-
Tapia, and A. Fock (2010). Crustal-scale structural architecture in thrust earthquakes and seismic swarms along the Chilean subduction
central Chile based on seismicity and surface geology: Implications zone (38°-18°S), Pure Appl. Geophys. 173, 1431–1449, doi: 10.1007/
for Andean mountain building, Tectonics 29, doi: 10.1029/ s00024-016-1280-5.
2009TC002480. Pacific Earthquake Engineering Research Center (PEER) (2019), Pacific
Giambiagi, L., A. Tassara, J. Mescua, M. Tunik, P. P. Alvarez, E. Godoy, G. Earthquake Engineering Research Center ground motion database,
Hoke, L. Pinto, S. Spagnotto, H. Porras, et al. (2015). Evolution of available at https://ngawest2.berkeley.edu (last accessed June 2019).
shallow and deep structures along the Maipo-Tunuyan transect (33° Pardo, M., and P. Acevedo (1984). Mecanismos de foco en la zona de Chile
40’s): From the Pacific coast to the Andean foreland, Geol. Soc. Central, Tralka 2, no. 3, 279–293.
Lond. Spec. Publ. 399, SP399-14, doi: 10.1144/SP399.14. Pérez, A., J. A. Ruiz, G. Vargas, R. Rauld, S. Rebolledo, and J. Campos
Giambiagi, L. B., and V. A. Ramos (2002). Structural evolution of the Andes (2014). Improving seismotectonics and seismic hazard assessment
between 33°30′ and 33°45′s above the transition zone between the flat along the San Ramón Fault at the eastern border of Santiago city,
and normal subduction segment, Argentina and Chile, J. South Am. Chile, Nat. Hazards 71, no. 1, 243–274, doi: 10.1007/s11069-013-
Earth Sci. 15, 101–116, doi: 10.1016/S0895-9811(02)00008-1. 0908-3.
Giambiagi, L. B., P. P. Alvarez, E. Godoy, and V. A. Ramos (2003). The Poiata, N., C. Satriano, J.-P. Vilotte, P. Bernard, and K. Obara (2016).
control of pre-existing extensional structures in the evolution of the Multiband array detection and location of seismic source recorded
southern sector of the Aconcagua fold and thrust belt, by dense seismic network, Geophys. J. Int. 205, 1548–1573, doi:
Tectonophysics 369, 1–19. 10.1093/gji/ggw071.
Giambiagi, L. B., V. A. Ramos, E. Godoy, P. P. Alvarez, and S. Orts (2003). Potin, B. (2016). Les Alpes Occidentales: Tomographie, localization de
Cenozoic deformation and tectonic style of the Andes, between 33° séismes et topographie du Moho, Ph.D. Dissertation, University
and 34° south latitude, Tectonics 22, no. 4, 1041, doi: 10.1029/ Grenoble-Alpes (France), ISTerre.
2001TC001354. Quigley, M. C., M. W. Hughes, B. A. Bradley, S. van Ballegooy, C. Reid, J.
Goldstein, P., D. Dodge, M. Firpo, and L. Minner (2003). SAC2000: Signal Morgenroth, T. Horton, B. Duffy, and J. R. Pettinga (2016). The 2010–
processing and analysis tools for seismologists and engineers, in The 2011 Canterbury earthquake sequence: Environmental effects, seismic
IASPEI International Handbook of Earthquake and Engineering triggering thresholds and geologic legacy, Tectonophysics 672/673,
Seismology, W. H. K. Lee, H. Kanamori, P. C. Jennings, and C. 228–274, doi: 10.1016/j.tecto.2016.01.044.
Kisslinger (Editors), Academic Press, London, United Kingdom. Reasenberg, P., and D. Oppenheimer (1985). FPFIT, FPPLOT and FPPAGE;
Graizer, V., and E. Kalkan (2016). Summary of the GK15 ground-motion Fortran computer programs for calculating and displaying earthquake
prediction equation for horizontal PGA and 5% damped PSA from fault-plane solutions, U.S. Geol. Surv. Open-File Rept. 85-739, doi:
shallow crustal continental earthquakes, Bull. Seismol. Soc. Am. 10.3133/ofr85739.
106, no. 2, 687–707, doi: 10.1785/0120150194. Riesner, M., R. Lacassin, M. Simoes, R. Armijo, R. Rauld, and G. Vargas
Hardebeck, J. L., and A. J. Michael (2006). Damped regional-scale stress (2017). Kinematics of the active west Andean fold-and-thrust belt
inversions: Methodology and example for southern California and the (Central Chile): Structure and long-term shortening rate, Tectonics
Coalinga aftershock sequence, J. Geophys. Res. 111, no. B11, doi: 36, 287–303, doi: 10.1002/2016TC004269.
10.1029/2005JB004144. Riesner, M., R. Lacassin, M. Simoes, D. Carrizo, and R. Armijo (2018).
Instituto Nacional de Prevención Sísmica (INPRES) (2019). Terremotos Revisiting the crustal structure and kinematics of the central Andes
históricos ocurridos en la República Argentina, available at http:// at 33.5°S: Implications for the mechanics of Andean Mountain build-
contenidos.inpres.gov.ar/sismologia/historicos (last accessed July 2019). ing, Tectonics 37, 1347–1375, doi: 10.1002/2017TC004513.
Ji, S., S. Sun, Q. Wang, and D. Marcotte (2010). Lamé parameters of Ruiz, S., and R. Madariaga (2018). Historical and recent large megathrust
common rocks in the Earth’s crust and upper mantle, J. Geophys. earthquakes in Chile, Tectonophysics 733, 37–56, doi: 10.1016/
Res. 115, no. B06314, doi: 10.1029/2009JB007134. j.tecto.2018.01.015.
Kanamori, H. (1977). The energy release in great earthquakes, J. Geophys. Ruiz, S., J.-B. Ammirati, F. Leyton, L. Cabrera, B. Potin, and R. Madariaga
Res. 82, no. 20, 2981–2987, doi: 10.1029/JB082i020p02981. (2019). The January 2019 (Mw 6.7) Coquimbo earthquake: Insights
Kanaori, Y., and S.-I. Kawakami (1996). The 1995 7.2 magnitude Kobe from a seismic sequence within the Nazca plate, Seismol. Res. Lett.
earthquake and the Arima-Takatsuki tectonic line: Implications of doi: 10.1785/0220190079.
the seismic risk for central Japan, Eng. Geol. 43, nos. 2/3, 135– Sepúlveda, S., M. Astroza, E. Kausel, J. Campos, E. Casas, S. Rebolledo,
150, doi: 10.1016/0013-7952(96)00056-7. and R. Verdugo (2008). New findings on the 1958 Las Melosas earth-
Leyton, F., S. A. Sepulveda, M. Astroza, S. Rebolledo, P. Acevedo, S. Ruiz, quake sequence, Central Chile: Implications for seismic hazard related
L. Gonzalez, and C. Foncea (2011). Seismic zonation of the Santiago to shallow crustal earthquake in subduction zones, J. Earthq. Eng. 12,
basin, Chile, 5th International Conference on Earthquake 432–455, doi: 10.1080/13632460701512951.
Geotechnical Engineering, Santiago, Chile. Shin, T, and T. Teng (2001). An overview of the 1999 Chi Chi, Taiwan, earth-
Leyton, P., A. Pérez, and J. Campos (2009). Anomalous seismicity in the quake, Bull. Seismol. Soc. Am. 91, no. 5, 895–913, doi: 10.1785/
lower crust of the Santiago basin, Chile, J. Phys. Earth Planet. In. 0120000738.
175, 17–25, doi: 10.1016/j.pepi.2008.03.016. Tarantola, A., and B. Valette (1982). Generalized nonlinear inverse problems
Lomax, A., J. Virieux, P. Volant, and C. Berge (2000). Probabilistic earth- solved using the least squares criterion, Rev. Geophys. 20, no. 2, 219–
quake location in 3D and layered models: Introduction of a Metropolis- 232, doi: 10.1029/RG020i002p00219.
Gibbs method and comparison with linear locations, in Advances in Thiele, R. (1980). Geología de la hoja Santiago, Región Metropolitana,
Seismic Event Location, C. H. Thurber and N. Rabinowitz (Editors), Carta Geológica de Chile, Inst. de Invest. Geol., Santiago, Chile, scale
Kluwer, Amsterdam, The Netherlands, 101–134. 1:250, 000, 51 pp.

Downloaded from https://pubs.geoscienceworld.org/ssa/bssa/article-pdf/109/5/1985/4833931/bssa-2019082.1.pdf


by Univ de Chile user
The Crustal Seismicity of the Western Andean Thrust (Central Chile, 33°–34° S) 1999

Vargas, G., Y. Klinger, T. K. S. Rockwell, L. Forman, S. Rebolledo, S. Baize, Jean-Baptiste Ammirati


R. Lacassin, and R. Armijo (2014). Probing large intraplate earth- Gabriel Vargas
quakes at the west flank of the Andes, Geology 42, no. 12, 1083– Sofía Rebolledo
Rachel Abrahami
1086, doi: 10.1130/G35741.1.
Facultad de Ciencias Físicas y Matemáticas
Vigny, C., A. Socquet, S. Peyrat, J. C. Ruegg, M. Métois, R. Madariaga, S.
Departamento de Geología
Morvan, M. Lancieri, R. Lacassin, J. Campos, et al. (2011). The
Universidad de Chile
2010 Mw 8.8 Maule megathrust earthquake of central Chile, Plaza Ercilla 803
monitored by GPS, Science 332, 1417–1421, doi: 10.1126/sci- 8320000 Santiago, Chile
ence.1204132. jbaptiste@ing.uchile.cl
Wallace, R. E. (1951). Geometry of shearing stress and relation to faulting, J.
Geol. 59, no. 2, 118–130. Bertrand Potin
Wells, D. L., and K. J. Coppersmith (1994). New empirical relation- Felipe Leyton
ships among magnitude, rupture length, rupture width, rupture area, Centro Sismológico Nacional
and surface displacement, Bull. Seismol. Soc. Am. 84, no. 4, 974– Universidad de Chile
1002. Blanco Encalada 2002
Wessel, P., and W. H. F. Smith (2006). New, improved version of generic 8320000 Santiago, Chile
mapping tools released, EOS Trans. AGU 79, no. 47, 579–579, doi:
10.1029/98EO00426. Sergio Ruiz
Yáñez, G., M. Muñoz, V. Flores-Aqueveque, and A. Bosch (2015). Gravity Facultad de Ciencias Físicas y Matemáticas
derived depth to basement in Santiago Basin, Chile: Implications for Departamento de Geofísica
its geological evolution, hydrogeology, low enthalpy geothermal, soil Universidad de Chile
Blanco Encalada 2002
characterization and geo-hazards, Andean Geol. 42, no. 2, 190–212,
8320000 Santiago, Chile
doi: 10.5027/andgeoV42n2-a01.
Zhang, L., J. Li, W. Liao, and Q. Wang (2016). Source rupture process of
the 2015 Gorkha, Nepal Mw7.9 earthquake and its tectonic implica-
tions, Geodes. Geodynam. 7, no. 2, 124–131, doi: 10.1016/ Manuscript received 1 April 2019;
j.geog.2016.03.001. Published Online 3 September 2019

Downloaded from https://pubs.geoscienceworld.org/ssa/bssa/article-pdf/109/5/1985/4833931/bssa-2019082.1.pdf


by Univ de Chile user

You might also like