ArticlePDF Available

Abstract and Figures

The subduction zone in northern Chile is a well-identified seismic gap that last ruptured in 1877. The moment magnitude (Mw) 8.1 Iquique earthquake of 1 April 2014 broke a highly coupled portion of this gap. To understand the seismicity preceding this event, we studied the location and mechanisms of the foreshocks and computed Global Positioning System (GPS) time series at stations located on shore. Seismicity off the coast of Iquique started to increase in January 2014. After 16 March, several Mw > 6 events occurred near the low-coupled zone. These events migrated northward for ~50 kilometers until the 1 April earthquake occurred. On 16 March, on-shore continuous GPS stations detected a westward motion that we model as a slow slip event situated in the same area where the mainshock occurred.
Content may be subject to copyright.
DOI: 10.1126/science.1256074
, 1165 (2014);345 Science
et al.S. Ruiz
8.1 earthquake
w
M
Intense foreshocks and a slow slip event preceded the 2014 Iquique
This copy is for your personal, non-commercial use only.
clicking here.colleagues, clients, or customers by , you can order high-quality copies for yourIf you wish to distribute this article to others
here.following the guidelines can be obtained byPermission to republish or repurpose articles or portions of articles
): September 5, 2014 www.sciencemag.org (this information is current as of
The following resources related to this article are available online at
http://www.sciencemag.org/content/345/6201/1165.full.html
version of this article at: including high-resolution figures, can be found in the onlineUpdated information and services,
http://www.sciencemag.org/content/suppl/2014/07/23/science.1256074.DC1.html
can be found at: Supporting Online Material
http://www.sciencemag.org/content/345/6201/1165.full.html#ref-list-1
, 11 of which can be accessed free:cites 27 articlesThis article
http://www.sciencemag.org/cgi/collection/geochem_phys
Geochemistry, Geophysics subject collections:This article appears in the following
registered trademark of AAAS. is aScience2014 by the American Association for the Advancement of Science; all rights reserved. The title CopyrightAmerican Association for the Advancement of Science, 1200 New York Avenue NW, Washington, DC 20005.
(print ISSN 0036-8075; online ISSN 1095-9203) is published weekly, except the last week in December, by theScience
on September 5, 2014www.sciencemag.orgDownloaded from on September 5, 2014www.sciencemag.orgDownloaded from on September 5, 2014www.sciencemag.orgDownloaded from on September 5, 2014www.sciencemag.orgDownloaded from on September 5, 2014www.sciencemag.orgDownloaded from on September 5, 2014www.sciencemag.orgDownloaded from
excluded for SN2014J (34). But in this case, the
accreted material is mostly He, and the accre-
tion rate can be very high, up to 10
4
solar masses
per year (44). If the white dwarf is rapidly ro-
tating or if mass is accreted faster than it loses
angular momentum and thus spreads over the
white dwarf, a He belt will be accumulated.
An equatorial ring as inferred here might not
be that uncommon. Recently, Hubble Space Tele-
scope imaging of the light echo from the re-
current nova T Pyx revealed a clumpy ring (45).
Once this belt becomes dense enough, explosive
He burning may be ignited, leaving an ejecta con-
figuration as shown in Fig. 3. This may be con-
sistent with the observed gamma-ray and optical
signals. Our radiation transfer simulations in UV/
optical/near-IR (fig. S10) show that the Ni belt
would not produce easily distinguishable features
but would result in a normal SN Ia appearance,
not only for a pole-on observer but also for an
equatorial observer. In view of this, the inter-
pretation of having this type of explosion as a
common scenario is not rejected by statistical
arguments (see the supplementary materials for
more details).
The evolution of the
56
Co gamma-ray signal
should reveal further aspects of the
56
Ni distri-
bution in SN2014J. These lines with associated
continua have been recognized to emerge in data
from both INTEGRAL instruments (46),asmore
of the total
56
Ni produced in the SN becomes
visible when the gamma-ray photosphere recedes
into the SN interior.
REFERENCES AND NOTES
1. J. Fossey, B. Cooke, G. Pollack, M. Wilde, T. Wright, Central
Bureau Electronic Telegrams No. 3792 (2014).
2. Y. Cao, M. M. Kasliwal, A. McKay, A. Bradley, Astronomers
Telegram, no. 5786 (2014).
3. W. Zheng et al., Astrophys. J. Lett. 783, L23 (2014).
4. J. J. Dalcanton et al., Astrophys. J. 183 (suppl.), 67108 (2009).
5. A. Goobar, B. Leibundgut, Annu. Rev. Nucl. Part. Sci. 61,
251279 (2011).
6. W. Hillebrandt, J. Niemeyer, Annu. Rev. Astron. Astrophys. 38,
191230 (2000).
7. W. Hillebrandt, M. Kromer, F. K. Röpke, A. J. Ruiter, Front. Phys.
8, 116143 (2013).
8. W. Li et al., Nature 480, 348350 (2011).
9. J. F. Guillochon, M. Dan, E. Ramirez-Ruiz, S. Rosswog,
Astrophys. J. 709, L64L69 (2010).
10. S. Rosswog, D. Kasen, J. Guillochon, E. Ramirez-Ruiz,
Astrophys. J. Lett. 705, L128L132 (2009).
11. M. Fink et al., Astron. Astrophys. 514, A53 (2010).
12. R. Pakmor et al., Astrophys. J. 747, L10 (2012).
13. F. K. Röpke et al., Astrophys. J. Lett. 750, L19 (2012).
14. M. C. P. Bours, S. Toonen, G. Nelemans, Astron. Astrophys.
552, A24 (2013).
15. P. Ruiz-Lapuente et al., Nature 431, 10691072 (2004).
16. W. E. Kerzendorf et al., Astrophys. J. 701,16651672 (2009).
17. B. E. Schaefer, A. Pagnotta, Nature 481, 164166 (2012).
18. J. S. Bloom et al., Astrophys. J. Lett. 744, L17 (2012).
19. B. J. Shappee, K. Z. Stanek, R. W. Pogge, P. M. Garnavich,
Astrophys. J. 762, L5 (2013).
20. A. Goobar et al., Astrophys. J. 784, L12 (2014).
21. P. Höflich, Nucl. Phys. A. 777, 579600 (2006).
22. S. A. Sim, P. A. Mazzali, Mon. Not. R. Astron. Soc. 385,
1681 (2008).
23. J. Isern et al., Astron. Astrophys. 552, A97 (2013).
24. A. Summa et al., Astron. Astrophys. 554, A67 (2013).
25. K. Maeda et al., Astrophys. J. 760, 54 (2012).
26. L.-S. The, A. Burrows, Astrophys. J. 786, 141 (2014).
27. C. Winkler et al., Astron. Astrophys. 411,L1L6 (2003).
28. E. Kuulkers, INTEGRAL Target of Opportunity observations
of the type Ia SN2014J in M82. Astronomers Telegram,
no. 5835 (2014).
29. G. Vedrenne et al., Astron. Astrophys. 411, L63L70 (2003).
30. J. P. Roques et al., Astron. Astrophys. 411,L91L100 (2003).
31. T. Siegert, thesis, Technische Universität München, Munich,
Germany (2013).
32. R. Diehl et al., Astron. Astrophys. 441, L117 (2003).
33. P. E. Nugent et al., Nature 480, 344347 (2011).
34. M. T. B. Nielsen, M. Gilfanov, A. Bogdan, T. E. Woods,
G. Nelemans, Upper limits on the luminosity of the progenitor
of type Ia supernova SN2014J. http://arxiv.org/abs/1402.2896
(2014).
35. P. L. Kelly, et al., http://arxiv.org/abs/1403.4250 (2014).
36. A. J. Ruiter, K. Belczynski, S. A. Sim, I. R. Seitenzahl,
D. Kwiatkowski, Mon. Not. R. Astron. Soc. 440,L101L105
(2014).
37. M. Fink, W. Hillebrandt, F. K. Röpke, Astron. Astrophys. 476,
11331143 (2007).
38. R. Moll, S. E. Woosley, Am. Astron. Soc. Abstr. 221, no. 253.22
(2013).
39. M. Kromer et al., Astrophys. J. 719, 10671082 (2010).
40. P. Nugent, E. Baron, D. Branch, A. Fisher, P. H. Hauschildt,
Astrophys. J. 485, 812819 (1997).
41. R. Kippenhahn, H. C. Thomas, Astron. Astrophys. 63, 265 (1978).
42. A. L. Piro, L. Bildsten, Astrophys. J. 603, 252264 (2004).
43. J. Greiner, in Lecture Notes in Physics (Springer Verlag, Berlin,
1996), vol. 472, pp. 299337.
44. W. Y. Law, H. Ritter, Astron. Astrophys.123,3338 (1983).
45. J. L. Sokoloski, A. P. S. Crotts, S. Lawrence, H. Uthas,
Astrophys. J. 770, L33 (2013).
ACKNOW LEDGM ENTS
This research was supported by the Deutsche Forschungsgemeinschaft
(DFG) cluster of excellence Origin and Structure of the Universe
and from DFG Transregio Project No. 33 Dark Universe.
S.A.G. acknowledges support from the Russian Academy of
Sciences, program RAS P-21. F.K.R. was supported by the DFG
(Emmy Noether Programm RO3676/1-1) and the ARCHES
prize of the German Ministry for Education and Research. The
work by K.M. is partly supported by a Japan Society for the
Promotion of Science Grant-in-Aid for Scientific Research (grant
no. 23740141 and 26800100) and WPI, at the Ministry of
Education, Culture, Sports. Science & Technology. We are grateful
to E. Kuulkers for handling the observations and to X. Zhang for
preparing our SPI data for the INTEGRAL SN2014J campaign.
The INTEGRAL/SPI project has been completed under the
responsibility and leadership of CNES Toulouse. We are grateful to
the agencies and institutions of ASI, CEA, CNES, DLR, ESA, INTA,
NASA, and OSTC for support of this ESA space science mission.
INTEGRALs data archive (http://www.isdc.unige.ch/integral/
archive#DataRelease) is at the INTEGRAL Science Data Center
in Versoix, Switzerland, and includes the SN2014J data used
in this paper.
SUPPLEMENTARY MATERIALS
www.sciencemag.org/content/345/6201/1162/suppl/DC1
Materials and Methods
Supplementary Text
Figs. S1 to S10
Tables S1 and S2
References (4658)
14 April 2014; accepted 21 July 2014
Published online 31 July 2014;
10.1126/science.1254738
EARTHQUAKE DYNAMICS
Intense foreshocks and a slow
slip event preceded the 2014
Iquique M
w
8.1 earthquake
S. Ruiz,
1
*M. Metois,
2
A. Fuenzalida,
3
J. Ruiz,
1
F. Leyton,
4
R. Grandin,
5
C. Vigny,
6
R. Madariaga,
6
J. Campos
1
The subduction zone in northern Chile is a well-identified seismic gap that last ruptured
in 1877. The moment magnitude (M
w
) 8.1 Iquique earthquake of 1 April 2014 broke a
highly coupled portion of this gap. To understand the seismicity preceding this event,
we studied the location and mechanisms of the foreshocks and computed Global
Positioning System (GPS) time series at stations located on shore. Seismicity off the coast
of Iquique started to increase in January 2014. After 16 March, several M
w
> 6 events
occurred near the low-coupled zone. These events migrated northward for ~50 kilometers
until the 1 April earthquake occurred. On 16 March, on-shore continuous GPS stations
detected a westward motion that we model as a slow slip event situated in the same
area where the mainshock occurred.
Since the giant moment magnitude (M
w
)
8.8 megathrust earthquake of 1877 (14),
the 500-km-long region stretching from
Arica (18°S) to the Mejillones Peninsula of
Chile (24.5°S) (Fig. 1) has had relatively few
major seismic events, with only moderate M
w
<8
events in 1933, 1967, and 2007 (49). On the basis
of recent geodetic data (5,10,11), the degree of
interseismic coupling (that is, the ratio between
theinterseismicsliprateandtheplate-convergence
velocity) in the area shows two distinct highly
coupled segments (Loa and Camarones) separated
by a low-coupling zone (LCZ) off the coast of Iquique
(5). On 1 April 2014, a ~150-km-long portion of
thegapbrokeinaM
w
8.1 earthquake after a strong
precursory activity that started on 16 March (12,13).
Understanding the complex nucleation phase
SCIENCE sciencemag.org 5 SEPTEMBER 2014 VOL 345 ISSUE 6201 1165
1
Departamento de Geofísica, Facultad de Ciencias Físicas y
Matemáticas, Universidad de Chile, Santiago, Chile.
2
Istituto
Nazionale di Geofisica e Vulcanologia, Centro Nazionale
Terremoti, Rome, Italy.
3
School of Environmental Sciences,
University of Liverpool, Liverpool, UK.
4
Centro Sismológico
Nacional, Facultad de Ciencias Físicas y Matemáticas,
Universidad de Chile, Santiago, Chile.
5
Institut de Physique
du Globe de Paris, Sorbonne Paris Cité, Université Paris
Diderot, UMR 7154 CNRS, Paris, France.
6
Laboratoire de
Geologie, UMR 8538 CNRS Ecole Normale Superieure, Paris,
France.
*Corresponding author. E-mail: sruiz@dgf.uchile.cl
RESEARCH |REPORTS
preceding the M
w
8.1 mainshock and its coseismic
rupture should provide insights into the seismic
gap history, present-day seismic hazard, and nu-
cleation process of megathrust earthquakes.
Seismic activity preceding the Iquique earth-
quake initiated in a region between 19.5°S and
21°S, where coupling ranges from 0.2 to 0.5 (5);
that is, the plates are slowly creeping past each
other at a fraction of the plate rate (when the
plates are fully locked, coupling is 1.0). This
region was actively monitored because its seis-
mic activity had steadily increased since 2008,
1166 5SEPTEMBER2014VOL 345 ISSUE 6201 sciencemag.org SCIENCE
Fig. 1. Northern Chile seismic gap. Interseismic coupling was calculated from GPS measurements acquired in the zone since 2008. Slip contours are shown
with continuous lines for the 2007 M
w
7.7 Tocopilla, 1995 M
w
8.0 Antofagasta, and 2014 M
w
8.1 and 7.6 Iquique earthquakes. Precursory seismicity from January to
March 2014 is shown with light blue circles. Foreshocks from March 2014 are shown with open squares.The dashed lines denote the surface projection of the
subduction interface isodepths.(Inset) Normalized probability that two or more M
w
< 7 earthquakes occurred since 2008 (CSN catalog) as a function of the value
of interseismic coupling. The upper right corner indicates the mean and standard deviation of the best-fit normal distribution.
RESEARCH |REPORTS
with repeated interplate thrust events of magni-
tude <4.0. Several seismic clusters were located
in this region by the National Seismological Cen-
terofChile(CSN)(Fig.1andfig.S1);someof
these clusters were associated with persistent mi-
croseismicity observed by the nearest seismolog-
ical stations (fig. S2). Global catalogs reveal an
increase in seismicity near Iquique since 2005,
compared with the previous 10 years (fig. S3).
To understand the seismicity that preceded the
1 April mainshock, we used the events listed in
the CSN catalog to relocalize and estimate the
focal mechanism of several foreshocks of this
sequence. Simultaneously, we calculated the Glob-
al Positioning System (GPS) time series of the
closest permanent stations up to the date of
the mainshock and inverted for the slip rate on
the plate interface (14).
The most recent seismic activity in northern
Chile started on 4 January 2014, when a M
w
5.7
interplate thrust event took place at 20.69°S,
70.80°W on the southern edge of the Iquique
LCZ (5). On 8 January, another M
w
5.7 event
occurred in the same area. The automatic loca-
tion system of CSN identified 30 events in a
smaller region of 15 km by 15 km, in the period
from 4 to 24 January 2014. Then on 12 January,
a small cluster of six identified events occurred
north of the previous ones at (19.7°S, 71.0°W).
During February 2014, another small cluster oc-
curred near 19.4°S, 71.0°W, where 16 events with
local magnitude (ML) 2.4 to 4.0 were identified
(Fig. 1, fig. S4, and table S1). On 15 March, the
area was reactivated with 11 events of ML 2.6 to
4.6, followed on 16 March by the first big fore-
shock of M
w
6.7, ~50 km south and with a slightly
deeper centroid (fig. S4). This event triggered a
persistent precursory seismicity, including a
M
w
6.3 earthquake on 22 March, which slowly
moved northward and lasted until the occur-
rence of the 1 April 2014 M
w
8.1 event (Fig. 2 and
fig. S4).
Overall, these foreshocks delineate a region
that spans ~150 km along the strike of the sub-
duction zone. We relocated this precursory seis-
micity and computed regional seismic moment
tensor using a linear time-domain, broadband
waveform inverse method (15)(Fig.2).Thecen-
troid of the M
w
6.7 precursor of 16 March was
only 10 km deep, in an area where the seismo-
genic interface is at a depth of ~20 km (16)(Fig.2
and fig. S5). This event had a reverse focal mecha-
nism with a strike of 277° that is at a sharp angle
with respect to the trench (Fig. 2A). Over the
course of 1 week, a persistent seismicity occurred
in a zone of 10-km radius (Fig. 2C); most of these
events were located inside the shallow South
American plate and had very diverse focal mecha-
nisms. On 22 March, a new M
w
6.3 foreshock
occurred ~30 km north of the 16 March foreshock.
After this event, precursory seismicity moved to
the vicinity of the 22 March foreshock with depths
and mechanisms indicatingthatitoccurredat
the plate interface (Fig. 2B).
From 16 March until the 1 April mainshock
(i.e., for 17 days), all of the continuous GPS (cGPS)
stations located along the coast between Iquique
and Pisagua started to move trenchward (Fig. 3).
This slowly increasing westward motion is in
contrast to the usual inland-directed interseis-
mic motion. The displacements measured dur-
ing this period were quite large (>5 mm for the
stations located between Iquique and Pisagua
and ~1 cm at the PSGA station). Only a fraction of
this cumulative displacement (up to 20%) can be
attributed to the largest foreshock of 16 March
(M
w
6.7), suggesting that slow aseismic slip was
taking place offshore, concurrently with the de-
velopment of the precursory seismicity. Whether
this motion started slightly before or coincided
with the 16 March foreshock is beyond the cur-
rent GPS resolution. We inverted for the slip
distributions on the subduction interface that
best reproduce the observed displacements using
Okadas formulas for an elastic half-space (14)
(Fig. 4). We found a slip of ~0.8 m for the M
w
6.7 events of 16 March, located in a narrow area
SCIENCE sciencemag.org 5 SEPTEMBER 2014 VOL 345 ISSUE 6201 1167
Fig. 2. Seismicity preceding the Iquique earthquake. (A) Gray dots show the foreshocks from 16 to 31 March; the intensity of the gray color indicates the
depths of the events.The slip distribution of the M
w
8.1 and M
w
7.6 earthquakes inverted from far-field broadband records of the International Federation of Digital
Seismograph Networks is shown with the color. (B) A 15-km-wide cross section along the A-Aline shown in (A). (C) A 15-km-wide cross section along the B-Bline
shown in (A). In the vertical cross sections, we plot the focal mechanisms of events with M
w
> 4.6. Mechanisms were computed by broadband moment tensor
analysis. The gray curve shows the seismogenic contact according to (16).
RESEARCH |REPORTS
in the vicinity of the CSN epicenter. We then com-
puted the aseismic slip for the period from 10
March to the mainshock that extends over an
area of 70 km by 20 km located between the fore-
shocks and the coast (Fig. 4).
The 1 April 2014 Iquique earthquake started
with a small shock at the northern end of the
region activated by the precursors that occurred
in March (19.57°S, 70.91°W). The peak seismic
moment release rate during the M
w
8.1 earth-
quake took place ~30 s after the initial nucleation
(fig. S6). We used standard teleseismic methods
(17) to invert for the coseismic slip of the main
event and its large M
w
7.6aftershockthatoc-
curred 2 days later (Fig. 2 and figs. S6 and S7).
The maximum slip associated with these events
was deeper than that of the March precursors
and ~20 km inland, affecting areas of higher
coupling (>0.6). Coseismic slip appears to over-
lap (at least partially) with the slow slip event
(SSE) (Fig. 4), but no substantial coseismic slip
occurred in the LCZ (Fig. 1). Whereas precursory
seismicity migrated northward, the mainshock
and its aftershock propagated toward the south,
similar to what was observed for the 2007 M
w
7.8
Tocopilla earthquake (7,8).
TheLCZoffthecoastofIquiquecanbeas-
sumed to play a key role in the events that
occurred in northern Chile during, and in the
20 years preceding, this precursory sequence.
The seismic swarms detected since 2008 oc-
curred on the edges of the LCZ, and most mod-
erate seismicity took place preferentially in zones
of intermediate coupling (see inset in Fig. 1 and
fig. S8), suggesting that aseismic creep occur-
ring on the LCZ triggered seismic activity in its
vicinity. Up to now, SSEs have remained unde-
tected by the cGPS network operating in north-
ern Chile for more than a decade. Nevertheless,
the very inference of a LCZ off the coast of Iquique
implies some degree of accommodation of plate
convergence by aseismic slip, which might op-
erate by repeated SSEs. We postulate that until
now, either the magnitude of these SSEs was too
small, or they occurred too far from the coast
to be detected by GPS measurements. The only
major change we detected in plate convergence
in the area before 2014 was a long-term velocity
change at the cGPS station operating in Iquique
since 1995 (UAPE):its eastward velocity decreased
after 2005 by ~20%, from 19.5 to 15.2 mm/year
(fig. S9). This suggests that interseismic loading
has been decreasing in the Iquique area during the
past decade, probably reflecting a very SSE oc-
curring on the decadal scale. This change could
have been triggered by the deep intraslab 2005
M
w
7.7 Tarapacá earthquake that generated
little postseismic relaxation (fig. S9).
Together with the fact that the M
w
8.1 main-
shock and M
w
7.6aftershockrupturesdidnot
penetrate into the LCZ, the foreshock sequence
and slow slip preceding the M
w
8.1 event argue
for a creeping Iquique LCZ. Because SSEs are
often associated with seismic swarms (18), we
propose that the seismicity observed in northern
Chile since 2008 was triggered by a SSE, devel-
oping for several years and accelerating during
the final foreshock sequence as in the preslip
model of nucleation (19). As suggested by many
laboratory experiments (20), the SSE might have
occurred in the nucleation zone of the impend-
ing megathrust rupture (Fig. 4). This precursory
sequence included several shallow crustal events
that took place near the 16 March foreshock;
these events may be associated with the activation
of a listric fault in the outermost fore-arc. This
area is poorly known due to the lack of marine
seismic profiles, but it may be similar to the
eroded wedge enhanced by fracturing imaged
at 22°S (21).
Several other subduction earthquakes were
preceded by precursory seismic activity (12); in
particular, the 1985 Valparaiso M
w
8.0 (22)and
2010 Maule M
w
8.8 (23) Chilean events and the
1168 5SEPTEMBER2014VOL 345 ISSUE 6201 sciencemag.org SCIENCE
Fig. 3. Motion of coastal GPS stations preced-
ing the Iquique earthquake. (A)Northand(B)
east components relative to a linear evolution
model with seasonal variations estimated since
2012 (14).Thethickredlinedenotestheorigintime
of the mainshock, whereas the black dotted lines
show the occurrence times of the M
w
>6fore-
shocks. Error bars indicate 1sformal uncertainty.
RESEARCH |REPORTS
2011 Tohoku-Oki earthquake (24). However, the
size and duration of the precursory events of
the 1 April mainshock are distinct. The occurrence
of a SSE before the 2011 M
w
9.0 Tohoku-Oki
earthquake and possibly the 2001 M
w
8.4 Arequipa
earthquake (24,25) implies that SSE may be a
common precursory feature and a potential trig-
gering mechanism for large megathrust ruptures.
On the other hand, whether the shallowest part
of the Camarones segment is still coupled and
capableofproducingalargeearthquakeorcreep-
ing aseismically is beyond the current resolution
of the GPS network. The large aftershock of 3
April left the highly coupled Loa segment south
of Iquique largely untouched. Several earthquakes
of equivalent or larger magnitude may still rup-
ture the deep intermediate-coupling areas of
this segment.
REFERENCES AND NOTES
1. F. Montessus de Ballore, Historia Sísmica de los Andes
Meridionales (Editorial Cervantes, Santiago, 1916).
2. E. Kausel, Bol. Acad. Chil. Ciencias 3,812 (1986).
3. C. Lomnitz, Geofis. Panamericana 1, 151178 (1971).
4. D. Comte, M. Pardo, Nat. Hazards 4,2344 (1991).
5. M. Metois et al., Geophys. J. Int. 194,12831294
(2013).
6. M. Malgrange, R. Madariaga, Geophys. J. R. Astron. Soc.
73, 489505 (1983).
7. S. Peyr at et al., Geophys. J. Int. 182, 14111430
(2010).
8. A. Fuenzalida, B. Schurr, M. Lancieri, M. Sobiesiak,
R. Madariaga, Geophys. J. Int. 194,12161238
(2013).
9. E. R. Engdahl, A. Villaseñor, in International Handbook of
Earthquake and Engineering Seismology (Academic Press,
San Diego, 2002), part A, chap. 41.
10. M. Chlieh et al., J. Geophys. Res. 116,B12405
(2011).
11. M. Béjar-Pizarr o et al., Nat. Geosci. 6, 462467
(2013).
12. Th. Lay, H. Yue, E. Brodsky, C. An, Geophys. Res. Lett. 41,
38183825 (2014).
13. Y. Yagi et al., Geophys. Res. Lett. 41, 42014206
(2014).
14. Materials and methods are available as supplementary
materials on Science Online.
15. M. E. Pasyanos, D. S. Dreger, B. Romanowicz, Bull. Seismol.
Soc. Am. 86, 12551269 (1996).
16. G. Hayes, D. J. Wald, R. L. Johnson, J. Geophys. Res. 117,
B01302 (2012).
17. M. Kikuchi, H. Kanamori, Bull. Seismol. Soc. Am. 81,
23352350 (1991).
18. S. Ozawa, H. Suito, M. Tobita, Earth Planets Space 59,
12411245 (2007).
19. W. L. Ellsworth, G. C. Beroza, Science 268, 851855
(1995).
20.S.Latour,A.Schubnel,S.Nielsen,R.Madariaga,
S. Vinciguerra, Geophys. Res. Lett. 40,50645069
(2013).
21. E. Contreras-Reyes, J. Jara, I. Grevemayer, S. Ruiz, D. Carrizo,
Nat. Geosci. 5, 342345 (2012).
22. D. Comte et al., Science 233, 449453 (1986).
23. R. Madariaga, M. Métois, C. Vigny, J. Campos, Science 328,
181182 (2010).
24. A. Kato et al., Science 335, 705708 (2012).
25. J. C. Ruegg, M. Olcay, D. Lazo, Seismol. Res. Lett. 72,
673678 (2001).
ACKNO WLED GME NTS
We thank J. C. Ruegg for his initiative to study northern Chile
with GPS, followed by J. B. De Chabalier, A. Socquet, D. Carrizo,
and others after him. S.R., J.R., R.M., and J.C. acknowledge
the support of the Chilean National Science Foundation
project FONDECYT no.1130636 and S.R. of project FONDECYT
no.11130230. This work received partial support from
ANR-2011-BS56-017 and ANR-2012-BS06-004 of the French
Agence Nationale de la Recherche. This is Institut de Physique
du Globe de Paris contribution no. 3551. We thank Incorporated
Research Institutions for Seismology Data Management Center
(www.iris.edu/data/), International Plate Boundary Observatory
Chile (http://geofon.gfz-potsdam.de/waveform/), Centro
Sismológico Nacional (www.sismologia.cl), and International
Associated Laboratory Montessus de Ballore for making raw
data available to us.
SUPPLEMENTARY MATERIALS
www.sciencemag.org/content/345/6201/1165/suppl/DC1
Materials and Methods
Figs. S1 to S13
Tables S1 and S2
References (2632)
14 May 2014; accepted 15 July 2014
Published online 24 July 2014;
10.1126/science.1256074
SCIENCE sciencemag.org 5SEPTEMBER2014VOL 345 ISSUE 6201 1169
Fig. 4. Slip distribution preceding
the Iquique earthquake. The slip
distributions were inverted using
Okadas equations for an elastic
half-space from the surface
displacements observed during
the preseismic phase (14) (Fig. 3).
(A) Coseismic slip due to the 16
March M
w
6.7 earthquake.
(B) Aseismic slip for the period
ranging from 10 to 31 March 2014.
(C) Cumulative slip model for
displacements observed from
10 to 31 March 2014. Black and
green contours are the slip
distributions for the mainshock
and the main aftershock,
respectively, already presented in
Fig. 2. Mo, seismic moment.
RESEARCH |REPORTS
... This along-dip segmentation differs from one subduction zone to the other (Nishikawa et al., 2019) and we note more occurrences of SSEs along young, warm subduction zones (i.e., Nankai, Mexico, Cascadia), than old and cold ones. Finally, slow slip appears to be an important ingredient of the preparation phase of earthquakes (e.g., Ruegg et al., 2001;Ruiz et al., 2014;Radiguet et al., 2016;Socquet et al., 2017;Voss et al., 2018). More recently, it has been proposed that a significant fraction of observed geodetic displacement in seismically active regions results from the occurrence of slow slip events (Jolivet and Frank, 2020, and reference therein), suggesting a burst-like, episodic behavior of aseismic slip at all time scales from seconds to decades in places as varied as Mexico (Frank, 2016;Rousset et al., 2017;Frank and Brodsky, 2019), Cascadia (Michel et al., 2019a;Ducellier et al., 2022;Itoh et al., 2022), along the San Andreas Fault (Khoshmanesh and Shirzaei, 2018;Rousset et al., 2019;Michel et al., 2022), the Haiyuan fault in Tibet (Jolivet et al., 2015a;Li et al., 2021), on the Alto Tiberina and Pollino fault systems in Italy (Gualandi et al., 2017;Cheloni et al., 2017;Essing and Poli, 2022), or Japan (Nishimura et al., 2013;Takagi et al., 2019;Nishikawa et al., 2019;Uchida et al., 2020). ...
... Regular slow slip events have been documented mainly along warm subduction zones such as Cascadia, Nankai (southwest Japan), Mexico, or New Zealand (e.g., Graham et al., 2016;Nishikawa et al., 2019;Wallace, 2020;Michel et al., 2022, and references therein). Instead, observations of slow slip events in cold subduction zones such as off-shore Japan or Chile are sparse or indirect, through seismic swarms, repeaters, or slow earthquakes (Kato et al., 2012;Kato and Nakagawa, 2014;Gardonio et al., 2018;Nishikawa et al., 2019), and rarely with geodetic observations (Hino et al., 2014;Ruiz et al., 2014;Socquet et al., 2017;Boudin et al., 2021). Geodetic displacement corresponding to such slow slip events are usually of mm to cm-scale amplitude and require the development of novel and systematic methods to extract SSEs from noisy time series of geodetic data (Frank, 2016;Rousset et al., 2017;Michel et al., 2019a;Uchida et al., 2020;Itoh et al., 2022). ...
... Afterslip has been reported following large earthquakes, including the 1995 M w 8.1 Antofagasta (Chlieh et al., 2004;Pritchard and Simons, 2006), the 2001 M w 8.1 Arequipa (Ruegg et al., 2001;Melbourne, 2002), the 2007 M w 8.0 Pisco Remy et al., 2016) Figure 1). Geodetic transients interpreted as the signature of aseismic slip occurred in the days to months preceding the M w 8.4 Arequipa earthquake in 2001, before one of its largest aftershock, and preceding the Iquique earthquake in 2014 (e.g., Ruegg et al., 2001;Melbourne, 2002;Ruiz et al., 2014;Schurr et al., 2014;Socquet et al., 2017). Aseismic slip is considered responsible for a significant fraction of such geodetic transients (Twardzik et al., 2022). ...
Article
Full-text available
Detections of slow slip events (SSEs) are now common along most plate boundary fault systems at the global scale. However, no such event has been described in the south Peru - north Chile subduction zone so far, except for the early preparatory phase of the 2014 Iquique earthquake. We use geodetic template matching on GNSS-derived time series of surface motion in Northern Chile to extract SSEs hidden within the geodetic noise. We detect 33 events with durations ranging from 9 to 40 days and magnitudes from Mw 5.6 to 6.2. The moment released by these aseismic events seems to scale with the cube of their duration, suggesting a dynamic comparable to that of earthquakes. We compare the distribution of SSEs with the distribution of coupling along the megathrust derived using Bayesian inference on GNSS- and InSAR-derived interseismic velocities. From this comparison, we obtain that most SSEs occur in regions of intermediate coupling where the megathrust transitions from locked to creeping or where geometrical complexities of the interplate region have been proposed. We finally discuss the potential role of fluids as a triggering mechanism for SSEs in the area.
... Before the 2007 Tocopilla earthquake there was a marked deficit followed by a period of surplus covering the entire year 2008. Subsequently, the compound process grows during 2009 taking off in 2010 and 2011 to undergo another takeoff event during 2012 and 2013 that starts a period of high deformation associated with the preparation phase of the 2014 Iquique earthquake (Ruiz et al., 2014) which, according to this graph, lasts 2 months. After the 2014 Iquique earthquake there is a period in which there is a balance with respect to the linear tectonic load. ...
... Subsequently there is a build-up stage at 23 mmy −1 rate (Φ ≃ 24%) culminating in two important dates. The July 10, 2013 6.1 event at 69.5 • W, 19.4 • S, 116 km depth and the January 4, 2014 5.7 already reported by Ruiz et al. (2014) as the start of the 2014 Iquique earthquake preparation phase. After the 2014 earthquake, equilibrium is reached with a relaxation period equal to the 2007 Tocopilla earthquake ending in August 2014 that is 5 months. ...
... The intersegment has not suffered earthquakes of magnitude greater than 7.7 since 1877 (Vigny and Klein, 2022), additionally in Figure 8 it is possible to appreciate the epicenters of events with magnitude greater than 7.7 recently recorded in the global catalog (U. S. Geological Survey, 2021), it can be seen that the rim suffered an aftershock in 2014 (Ruiz et al., 2014) around the 30 km isobath and that the 2007 Tocopilla poearthquake is at the other extreme. In this intersegment Ruiz et al. (2014) infers the occurrence of a slow earthquake prior to the 2014 Iquique earthquake, according to our analysis, here the regional balance process would be consistent with slow deformation. In the elongated strip towards the Cordillera between 18 • S and 24 • S there is also a seismicity deficit. ...
Preprint
Full-text available
The seismic cycle in subduction zones comprehends a phenomena of build-up and release of strain, which is punctuated by the occurrence of earthquakes. Nonetheless, the occurrence of earthquakes themselves depends on the relative plate velocity and on lateral heterogeneities that ponderate the energy release. This characteristic is exploited in order to obtain a seismic cycle representation in Northern Chile, using data from the IPOC catalog in the years 2007–2014. We propose and evaluate a scaling relationship for the energy released by earthquakes in a determined scale, depending on the elastic modulus, earthquake displacement and mean stress drop. Displacement, on the other hand, is obtained assuming that the seismicity rate is locally homogeneous and that the averaged regional balance process, which counters tectonic displacement in time due to plates relative velocity with the cumulative sum of earthquake displacements, washes out on the long term. This framework allowed us to obtain a seismic cycle representation between megathrust earthquakes from 2007 and 2014, accounting for a variety of phenomena observed.
... The relationship between SSEs and earthquake hazards remains an important topic of active research, with some observations suggesting SSEs preceding several great subduction events, such as the 2011 M W 9.0 Tohoku-Oki and 2014 M W 8.1 Iquique earthquakes (Ito et al., 2013;Kato et al., 2012;Ruiz et al., 2014), as well as earthquake ruptures penetrating into regions known to host slow slip (Lin et al., 2020). A variety of fault models based on low-velocity, laboratory-derived rate-and-state friction laws (e.g. ...
Article
Full-text available
Plain Language Summary Earthquakes are thought to predominantly occur along sections of faults that appear stuck and actively accumulating strain under tectonic plate motion. Other fault regions observed to be slowly slipping are thought to release some of this strain without causing strong shaking, potentially limiting the location and amount of fault slip in earthquakes. Here we present numerical simulations of long‐term fault slip that add to a body of work showing how fault areas can host different styles of slow slip for several decades prior to failing destructively when pushed by an incoming earthquake rupture. Our models show how relatively short‐term observations of slow fault slip compared to the recurrence of large earthquakes over several centuries can mask fault regions that are capable of experiencing substantial slip in future earthquakes. Importantly, our simulations suggest that if a fault region is capable of failing during an earthquake, then observations of slow slip may indicate that the region is favorably stressed to fail in a future earthquake, representing a qualitatively different interpretation of slow slip for seismic hazard.
... Irregular initial rupture phases have been observed for other earthquakes (e.g., Abercrombie & Mori, 1994;Ellsworth & Beroza, 1995), although the 10-s duration of our E0 episode is longer than most cases documented in the literature. One exception is the 2014 M W 8.1 Iquique, Chile, earthquake, which had a long (20 s) initial rupture phase (e.g., Yagi et al., 2014) and was preceded by swarm-like foreshock activity that lasted around 2 weeks (e.g., Kato & Nakagawa, 2014;Ruiz et al., 2014). These observations raise the question of what environment fosters the development of the type of rupture behavior observed in the 2024 Noto Peninsula earthquake. ...
Article
Full-text available
Plain Language Summary On 1 January 2024, a moment magnitude 7.5 earthquake occurred in the northern Noto Peninsula, Japan. The strong ground motion and tsunami associated with the earthquake caused severe damage to buildings and infrastructure, resulting in at least 245 causalities in the affected areas. The Noto Peninsula is affected by northwest‐southeast compression, and active reverse faults are known along the northern coast of the peninsula and its offshore region. Before the 2024 earthquake, the source region experienced long‐lasting earthquake swarm activity, which is a set of seismic events without an obvious mainshock‐aftershock pattern. Our seismological analysis found that there was a 10‐s‐long initial rupture episode around the hypocenter that overlapped with the earthquake swarm region. The initial rupture was followed by a series of three different rupture episodes on differently oriented fault segments. This earthquake highlights a multi‐scale rupture growth across a segmented fault network after a very quiet initial rupture process that was controlled by the preceding earthquake swarms and associated aseismic deformation related to fluid injection from depth. The rupture process advances our understanding of earthquake source physics and can lead to a better assessment of future earthquake hazards.
... About 1,500 km further north, the Mw8.1 Iquique earthquake occurred on 1 April 2014, at a shallow depth of 23 km . The thrust faulting slipped at least ∼150 km along the primary plate boundary interface between ∼19°and ∼21°S to generate a local tsunami (Hayes et al., 2014;Ruiz et al., 2014;Schurr et al., 2014;Soto et al., 2019). About 500 km north of the Maule earthquake location, the Mw8.3 Illapel earthquake happened at a depth of 22 km on 16 September 2015 (Lee et al., 2016). ...
Article
Full-text available
We developed an enhanced Kalman‐based approach to quantify abrupt changes and significant non‐linearity in vertical land motion (VLM) along the coast of Chile and the Antarctic Peninsula using a combination of multi‐mission satellite altimetry (ALT), tide gauge (TG), and GPS data starting from the early 1990s. The data reveal the spatial variability of co‐seismic and post‐seismic subsidence at TGs along the Chilean subduction zone in response to the Mw8.8 Maule 2010, Mw8.1 Iquique 2014, and Mw8.3 Illapel 2015 earthquakes that are not retrievable from the interpolation of sparse GPS observations across space and time. In the Antarctic Peninsula, where continuous GPS data do not commence until ∼1998, the approach provides new insight into the ∼2002 change in VLM at the TGs of +5.3 ± 2.2 mm/yr (Palmer) and +3.5 ± 2.8 mm/yr (Vernadsky) due to the onset of ice‐mass loss following the Larsen‐B Ice Shelf breakup. We used these data to constrain viscoelastic Earth model parameters for the northern Antarctic Peninsula, obtaining a preferred lithosphere thickness of 115 km and upper mantle viscosity of 0.9 × 10¹⁸ Pa s. Our estimates of regionally‐correlated ALT systematic errors are small, typically between ∼±0.5–2.5 mm/yr over single‐mission time scales. These are consistent with competing orbit differences and the relative errors apparent in ALT crossovers. This study demonstrates that, with careful tuning, the ALT‐TG technique can provide improved temporal and spatial sampling of VLM, yielding new constraints on geodynamic models and assisting sea‐level change studies in otherwise data sparse regions and periods.
... Figure as such based on some less than 8 Mw events, Domain X, the southernmost domain, is dominated by the giant 210 events of Valdivia 1960, 9.3 Mw, and1737, 9.0 Mw. 211 Adequate seismic coverage is available since 1985 in Chile. In this period, six large earthquakes have been 212 recorded: Valparaiso 1985, 8.0 Mw (Comte et al., 1986;Mendoza et al., 1994); Antofagasta 1995, 8.0 Mw 213 , Delouis et al., 1997Pritchard et al., 2002 andChlieh et al., 2004); Tocopilla 2007, 7.8 Mw 214 (Schurr et al., 2012); Maule 2010, 8.8 Mw (Delouis et al., 2010;Lay et al., 2010;Vigny et al., 2011;Koper et 215 al., 2012;Ruiz et al., 2012;Moreno et al., 2012;Lorito et al., 2011;Lin et al., 2013;Yue et al., 2014); Iquique 216 2014, 8.2 Mw (Ruiz et al., 2014;Hayes et al., 2014;Schurr et al., 2014;Lay et al., 2014), andIllapel 2015, 8.3 217 Mw (Melgar et al., 2016;Heidarzadeh et al., 2016;Li et al.,2016;Lee et al., 2016;Satake and Heidarzadeh, 218 2017). Given the large size of the Valdivia 1960 earthquake (9.3 Mw), we also include slip estimates for this 219 event based on surface deformation data (Barrientos and Ward, 1990). ...
Preprint
Full-text available
Plate coupling play a fundamental role in the way in which seismic energy is released during the seismic cycle. This process includes quasi-instantaneous release during megathrust earthquakes and long-term creep. Both mechanisms can coexist in a given subducting margin, defining a seismotectonic segmentation in which seismically active segments are separated by zones in which ruptures stop, classified for simplicity as asperities and barrier, respectively. The spatiotemporal stability of this segmentation has been a matter of debate in the seismological community for decades. At this regard, we explore in this paper the potential role of the interaction between geological heterogeneities in the overriding plate and fluids released from the subducting slab towards the subduction channel. As a case study, we take the convergence between the Nazca and South American plates between 18°–40° S, given its relatively simple convergence style and the availability of a high-quality instrumental and historical record. We postulate that trans-lithospheric faults striking at a high angle with respect to the trench behave as large fluid sinks that create the appropriate conditions for the development of barriers and promote the growth of highly coupled asperity domains in their periphery. We tested this hypothesis against key short- and long-term observations in the study area, obtaining consistent results. If the spatial distribution of asperities is controlled by the geology of the overriding plate, seismic risk assessment could be established with better confidence.
... Slow slip events (SSEs) are another research target of DA (Hirahara and Nishikiori 2019;Diab-Montero et al. 2023). Many observational studies have inferred that SSEs occurred prior to megathrust earthquakes (e.g., Ito et al. 2013;Ruiz et al. 2014;Radiguet et al. 2016;Voss et al. 2018); therefore, monitoring and predicting the spatio-temporal evolution of SSEs is important, particularly for assessing future earthquake potential. Hirahara and Nishikiori (2019) (hereafter referred to as HN19) first introduced the use of DA for the fault slip estimation problem of SSEs focusing on long-term SSEs in the Bungo Channel, southwest Japan. ...
Article
Full-text available
Monitoring and predicting fault slip behaviors in subduction zones is essential for understanding earthquake cycles and assessing future earthquake potential. We developed a data assimilation method for fault slip monitoring and the short-term prediction of slow slip events, and applied to the 2010 Bungo Channel slow slip event in southwest Japan. The observed geodetic data were quantitatively explained using a physics-based model with data assimilation. We investigated short-term predictability by assimilating observation data within limited periods. Without prior constraints on fault slip style, observations solely during slip acceleration predicted the occurrence of a fast slip; however, the inclusion of slip deceleration data successfully predicted a slow transient slip. With prior constraints to exclude unstable slip, the assimilation of data after slow slip event occurrence also predicted a slow transient slip. This study provides a tool using data assimilation for fault slip monitoring and prediction based on real observation data. Graphical Abstract
Preprint
In a recent publication, we showed that a stack of all GPS time series recorded before Magnitude ≥ 7.0 earthquakes suggests that large earthquakes start with a precursory phase of accelerating slow slip (Bletery and Nocquet, 2023). While no peer-reviewed comment or publication has formally contradicted this result, informal discussion has emerged on various platforms. We present here the different elements of discussion and address them through a series of tests. In particular, it has been proposed that correcting GPS time series from network common-mode noise makes the signal vanish. We confirm this result, but we show that this common-mode filtering procedure may inadvertently remove an existing tectonic signal. Moreover, the analysis of past records indicate that the likelihood that common-mode noise produces the signal we observe is well below 1 %. Additionally, we find that the signal is maximum at the location of the impending earthquakes, and for a slip direction (rake angle) close to the one of the upcoming events. The collective outcomes of these tests make very unlikely that the signal solely arises from noise. Even though the results of our tests do not irrefutably demonstrate the existence of a precursory phase of slow slip, they do support its existence. We hope that this study will motivate further work by others to provide a definite answer to the question of the tectonic origin of the observed signal and confirm or refute that large earthquakes start with a precursory phase of slow slip.
Article
Full-text available
Plain Language Summary Understanding the earthquake nucleation process has direct implications for earthquake physics and seismic hazards. Specifically, identifying the geophysical processes within fault zones that precede and result in subsequent earthquakes has been of great interest to the earthquake science community. This study explores how the ratio of compressional (P) wave speed to shear (S) wave speed changes in both space and time and their correlations with the subsequent seismicity evolution, focusing on the 2019 Ridgecrest earthquake sequence in California. In most slipping areas of the earthquakes, the P‐wave to S‐wave speed ratios are relatively uniform. However, in places where faults end, cross each other, or change direction, we observe unusual values. We find high ratios near the three major earthquakes on a small scale. Additionally, the ratios change where the main earthquake (magnitude 7.1) initiated. The ratios are low between a moderate (magnitude 5.4) and the main (magnitude 7.1) earthquake and increase after the main earthquake, indicating the presence of over‐pressurized fluids near the earthquake source. The associated high pore pressure might have helped nucleate the Ridgecrest mainshock. Our findings show that these speed ratios can be highly sensitive to seismic activities and could help us better understand how earthquakes start.
Article
Full-text available
On April 1, 2014, a great (Mw 8.1) interplate thrust earthquake ruptured in the northern portion of the 1877 earthquake seismic gap in northern Chile. The sequence commenced on March 16, 2014 with a magnitude 6.7 thrust event, followed by thrust-faulting aftershocks that migrated northward ~40 km over two weeks to near the mainshock hypocenter. Guided by short-period teleseismic P wave back-projections and inversion of deep-water tsunami wave recordings, a finite-fault inversion of teleseismic P and SH waves using a geometry consistent with long-period seismic waves resolves a spatially compact large-slip (~2-6.7 m) zone located ~30 km down-dip and ~30 km along-strike south of the hypocenter; down-dip of the foreshock sequence. The mainshock seismic moment is 1.7 × 1021 Nm with a fault dip of 18°, radiated seismic energy of 4.5-8.4 × 1016 J, and static stress drop of ~2.5 MPa. Most of the 1877 gap remains unbroken and hazardous.
Article
Full-text available
We analysed archival data of Chandra pre-explosion observations of the position of SN 2014J in M82. No X-ray source at this position was detected in the data, and we calculated upper limits on the luminosities of the progenitor. These upper limits allow us to firmly rule out an unobscured supersoft X-ray source progenitor with a photospheric radius comparable to the radius of white dwarf near the Chandrasekhar mass (∼1.38 M⊙) and mass accretion rate in the interval where stable nuclear burning can occur. However, due to a relatively large hydrogen column density implied by optical observations of the supernova, we cannot exclude a supersoft source with lower temperatures, kT ≲ 70 eV. We find that the supernova is located in the centre of a large structure of soft diffuse emission, about 200 pc across. The mass, ∼3 × 104 M⊙ and short cooling time of the gas, τcool ∼ 8 Myr, suggest that it is a supernova-inflated superbubble, associated with the region of recent star formation. If SN 2014J is indeed located inside the bubble, it likely belongs to the prompt population of Type Ia supernovae, with a delay time as short as ∼50 Myr. Finally, we analysed the one existing post-supernova Chandra observation and placed upper limit of ∼(1–2) × 1037 erg s−1 on the X-ray luminosity of the supernova itself.
Article
Full-text available
Segmentation can influence the extent of earthquake rupture and event magnitude: large megathrust earthquakes result from total rupture of relatively continuous segments of the subduction interface. Segmentation is attributed to variations in the frictional properties of the seismogenic zone or to topographic features on the down-going plate. Structures in the overriding plate may also influence segmentation, but their importance has been dismissed. Here, we investigate the links between interface segmentation at the North Chile seismic gap and a crustal-scale fault structure in the overriding plate that forms a coastal scarp of about 1km in height. We use satellite interferometric synthetic aperture radar (InSAR) and Global Positioning System (GPS) data to measure interseismic surface deformation between 2003 and 2009 and compare the deformation with rupture extent during well-documented earthquakes. From these data we infer the degree of coupling and segmentation at depth. We find that along a 500-km-long segment, the base of the strongly coupled seismogenic zone correlates with the line of the surface coastal scarp and follows the outline of the Mejillones Peninsula. This correlation implies that large-scale structures in the overriding plate can influence the frictional properties of the seismogenic zone at depth. We therefore suggest that the occurrence of megathrust earthquakes in northern Chile is controlled by the surface structures that build Andean topography.
Article
We have developed a method that inverts seismic body waves to determine the mechanism and rupture pattern of earthquakes. The rupture pattern is represented as a sequence of subevents distributed on the fault plane. This method is an extension of our earlier method in which the subevent mechanisms were fixed. In the new method, the subevent mechanisms are determined from the data and are allowed to vary during the sequence. When subevent mechanisms are allowed to vary, however, the inversion often becomes unstable because of the complex trade-offs between the mechanism, the timing, and the location of the subevents. Many different subevent sequences can explain the same data equally well, and it is important to determine the range of allowable solutions. Some constraints must be imposed on the solution to stabilize the inversion. We have developed a procedure to explore the range of allowable solutions and appropriate constraints. In this procedure, a network of grid points is constructed on the τ - I plane, where τ and I are, respectively, the onset time and the distance from the epicenter of a subevent; the best-fit subevent is determined at all grid points. Then the correlation is computed between the synthetic waveform for each subevent and the observed waveform. The correlation as a function of τ and I and the best-fit mechanisms computed at each τ - I grid point depict the character of allowable solutions and facilitate a decision on the appropriate constraints to be imposed on the solution. The method is illustrated using the data for the 1976 Guatemala earthquake.
Chapter
Probabilistic earthquake location with non-linear, global search methods allows the use of 3D models and produces comprehensive uncertainty and resolution information represented by a probability density function over the unknown hypocentral parameters. We describe a probabilistic earthquake location methodology and introduce an efficient Metropolis-Gibbs, non-linear, global sampling algorithm to obtain such locations. Using synthetic travel times generated in a 3D model, we examine the locations and uncertainties given by an exhaustive grid-search and the Metropolis-Gibbs sampler using 3D and layered velocity models, and by a iterative, linear method in the layered model. We also investigate the relation of average station residuals to known static delays in the travel times, and the quality of the recovery of known focal mechanisms. With the 3D model and exact data, the location probability density functions obtained with the Metropolis-Gibbs method are nearly identical to those of the slower but exhaustive grid-search. The location PDFs can be large and irregular outside of a station network even for the case of exact data. With location in the 3D model and static shifts added to the data, there are systematic biases in the event locations. Locations using the layered model show that both linear and global methods give systematic biases in the event locations and that the error volumes do not include the “true” location — absolute event locations and errors are not recovered. The iterative, linear location method can fail for locations near sharp contrasts in velocity and outside of a network. Metropolis-Gibbs is a practical method to obtain complete, probabilistic locations for large numbers of events and for location in 3D models. It is only about 10 times slower than linearized methods but is stable for cases where linearized methods fail. The exhaustive grid-search method is about 1000 times slower than linearized methods but is useful for location of smaller number of events and to obtain accurate images of location probability density functions that may be highly-irregular.
Article
On 23 June 2001, at 20:33 UTC, a major earthquake occurred near the coast of southern Peru, about 190 km west of Arequipa (16.15°S, 73.40°W). A revised magnitude of 8.4 was computed for this earthquake by the United States Geological Survey, NEIC (see http://neic.usgs.gov/neis/bulletin/01_EVENTS/0l0623203313/010623203313.html for further information). The earthquake occurred along the west coast of Peru (Figure 1) at the boundary between the Nazca and South American Plates and resulted from thrust faulting along this boundary as the Nazca Plate subducts beneath the South American Plate. The two plates are converging toward each other at a rate of 78 mm/yr (DeMets et al. , 1990). Southwestern Peru has a history of very large earthquakes. The 23 June shock originated just southeast of the source of a magnitude 7.7 earthquake that occurred in November 1996. This area of the plate boundary has not recorded a major rupture since the 1868 earthquake of magnitude approximately 8.5–9 (Dorbath et al. , 1990; Comte and Pardo, 1991). The 23 June shock occurred in a previously identified seismic gap (Nishenko, 1985). The 1868 earthquake was more destructive than the 23 June earthquake and produced a tsunami that killed hundreds of people along the South American coast. The 23 June earthquake is very similar to the 1995 Mw 8.1 Antofagasta earthquake (Ruegg et al. , 1996), which ruptured a zone of about 180 × 70 km2 of the Chilean coast, along which no major historical earthquake has been recognized before. The city of Arequipa, located 190 km east of the main shock epicenter, experienced damage during the 23 June earthquake. At Arequipa several precise positioning systems are permanently operating, including continuous GPS station of the IGS, DORIS beacon, Satellite Laser Ranging. In this study we give a quick report about surface deformation …
Article
The rupture process of the 2014 Iquique, Chile earthquake is inverted from teleseismic P-wave data applying a novel formulation that takes into account the uncertainty of Green's function, which has been a major error source in waveform inversion. The estimated seismic moment is 1.5 × 1021 Nm (Mw =8.1), associated with a 140 km long and 140 km wide fault rupture along the plate interface. The source process is characterized by unilateral rupture propagation. During the first 20 s the dynamic rupture front propagated from the hypocenter to the large asperity located about 50 km southward, crossing a remarkably active foreshock area at high velocity (of about 3.0 km/s), but small and irregular seismic moment release rate. Our result may suggest that the 20 s long initial phase was influenced by the stress drop due to the foreshock activity near the mainshock hypocenter. Moreover, the 2-weeks long swarm-like foreshock activity migrating roughly at 5 km/day towards the mainshock hypocenter, and possibly associated slow-slip, contributed to the stress accumulation prior to the Mw 8.1 megaquake. The mainshock initial rupture phase might have triggered the rupture of the large asperity, which had large fracture energy.
Article
The recent supernova (SN) in M82 was detected on January 21, 2014 at the UCL Observatory (CBET #3792). Originally it was designated PSN J09554214+6940260; it has been identified as a type Ia SN (ATel #5786) and is referred to as SN2014J (CBET #3792).
Article
We observe the nucleation phase of in-plane ruptures in the laboratory. We show that the nucleation is composed of two distinct phases, a quasi-static and an acceleration stage, followed by dynamic propagation. We propose an empirical model which describes the rupture length evolution: The quasi-static phase is described by an exponential growth while the acceleration phase is described by an inverse power law of time. The transition from quasi-static to accelerating rupture is related to the critical nucleation length, which scales inversely with normal stress in accordance with theoretical predictions, and to a critical surfacic power, which may be an intrinsic property of the interface. Finally, we discuss these results in the frame of previous studies and propose a scaling up to natural earthquake dimensions.