Вы находитесь на странице: 1из 5

LETTERS

PUBLISHED ONLINE: 30 JANUARY 2011 | DOI: 10.1038/NGEO1073

Limited overlap between the seismic gap and


coseismic slip of the great 2010 Chile earthquake
S. Lorito1 *, F. Romano1 , S. Atzori1 , X. Tong2 , A. Avallone1 , J. McCloskey3 , M. Cocco1 , E. Boschi1
and A. Piatanesi1

The Mw 8.8 mega-thrust earthquake and tsunami that occurred 32


on 27 February 2010 offshore the Maule region, Chile, was
not unexpected. A clearly identified seismic gap1–13 existed in
an area where tectonic loading has been accumulating since 1985 8,000
the great 1835 earthquake14 . Here we jointly invert tsunami 0.75 Santiago
7,000
and geodetic data to derive a robust model for the coseismic

0.25
6,000
34
slip distribution and induced coseismic stress changes. We 5,000

0.85
compare these with past earthquakes and the preseismic
4,000
locking distribution13 , to assess if the Maule earthquake has ¬1
yr 1928
filled the seismic gap. We find that the main slip patch is 6.8 cm Constitucion
3,000

`
located to the north of the gap, overlapping the rupture Latitude (° S) 0.75
2,000

Altitude (m)
zone of the Mw 8.0 earthquake that occurred in 1928, with 36 1,000
1939
a secondary concentration of slip to the south. The seismic 0

1835
0.85
gap was only partially filled and a zone of high preseismic Concepcion ¬1,000

`
locking remains unbroken, inconsistent with the assumption
¬2,000
that distributions of seismic rupture might be correlated with

0.50
¬3,000
preseismic locking. Moreover, we conclude that increased Mw 8.8
stress on the unbroken patch may in turn have increased the 38 27 Feb 2010
¬4,000
1960
probability of another major to great earthquake there in the ¬5,000
near future. ¬6,000
On 27 February 2010 at 06:34:14 utc an earthquake of moment ¬7,000
magnitude Mw 8.8 occurred offshore the Maule region, Chile, some
0 100 ¬8,000
360 km southwest of Santiago, with epicentre at 36.12◦ S, 72.90◦ W (km)
(http://earthquake.usgs.gov/earthquakes/eqarchives/epic/). The 40
ensuing tsunami caused severe damage along the adjacent coasts, 76 74 72 70
with reported maximum wave heights of more than 10 m at Consti- Longitude (°W)
tución (http://www.ngdc.noaa.gov/hazard/tsu.shtml), and a devas-
tating inundation in Juan Fernández islands, some 600 km offshore. Figure 1 | Location map of the 2010 Maule earthquake and the seismic
The earthquake and tsunami claimed together more than 500 lives. gap. Geographic location of the 27 February 2010 Mw 8.8 Maule
The 2010 Maule earthquake ruptured the rapidly converging earthquake, with epicentre (red star) between Concepción and
(6.8 cm yr−1 ; ref. 10) interface between the Nazca and South Constitución in South Central Chile. The green and white ‘beach ball’ is the
American plates and was the largest mega-thrust earthquake United States Geological Survey centroid moment tensor. Yellow stars are
recorded along this segment since the Mw 9.5 1960 Chile shock the epicentres of 1928, 1939, 1960 and 1985 earthquakes, with their
(Fig. 1). The source zone of the Maule earthquake features approximate source zones4,7,16,19 , which are shaded for thrust inter-plate
heterogeneous but locally high plate coupling13 . Tectonic stress has events. The 1960 source zone includes both 21 May Mw 7.9 and 22 May
probably accumulated there continuously since the most recent Mw 9.5 earthquakes19 . Orange lines are contours of the preseismic locking
major (Mw ∼ 8.5; ref. 4) subduction earthquake, experienced and distribution13 . The segment that probably contains the source zone of the
described by Darwin during the voyage of the Beagle in 1835 1835 earthquake (indicated by the black dashed line with arrows) is the
(ref. 14), storing more than 12 m of slip where the interface is zone of the Darwin14 gap, where a major earthquake was expected12 . The
fully locked. The more recent Mw 7.9 1939 earthquake was in fact red line with triangles is the trench between the Nazca and South America
an intra-plate event1–4 , and the segment between the Mw 9.5 1960 Plates26 . The white arrow indicates the approximate convergence direction
earthquake to the south, and the Mw 8.0 (ref. 4) 1928 earthquake to of the plates along with its estimated velocity10 .
the north, then contains a well defined seismic gap1–13 , referred to
here as the ‘Darwin gap’. Tsunami observations provide an indirect measure of the seafloor
The 2010 Maule rupture nucleated close to the coast, producing displacement, whereas geodetic observations (Interferometric
a significant displacement both on the sea floor and onshore. Synthetic Aperture Radar (InSAR), Global Positioning System

1 IstitutoNazionale di Geofisica e Vulcanologia, Via di Vigna Murata 605, 00143 Rome, Italy, 2 Institute of Geophysics and Planetary Physics, Scripps
Institution of Oceanography, University of California, San Diego, La Jolla, California 92093-0225, USA, 3 Environmental Sciences Research Institute, School
of Environmental Sciences, University of Ulster, Coleraine BT52 1SA, Northern Ireland. *e-mail: stefano.lorito@ingv.it.

NATURE GEOSCIENCE | ADVANCE ONLINE PUBLICATION | www.nature.com/naturegeoscience 1


© 2011 Macmillan Publishers Limited. All rights reserved.
LETTERS NATURE GEOSCIENCE DOI: 10.1038/NGEO1073

32 are mapped in Supplementary Fig. S1, values in Supplementary


Table S1). The major slip patch is north of the epicentre between 36
and 34◦ S, with a longitudinally elongated area of some 200×80 km,
and a slip concentration centred on 35–35.5◦ S with peaks of
1985 18–19 m. The highest slip is confined to about 60–110 km down-dip
from the trench, corresponding to a depth of 25–40 km. The
southern slip patch, which measures some 130 × 60 km, is centred
34 on about 37.5◦ S and is characterized by lower overall slip, up
to about 9–10 m at a similar depth to the northern patch. The
slip direction is grossly consistent with plate convergence, with
1 1928 a dominant thrust and minor right-lateral components. Our slip
yr¬
6.8 cm model corresponds to a seismic moment of 1.55 × 1022 N m and
magnitude Mw = 8.8 using a rigidity value of 30 GPa.
Latitude (° S)

The combined tsunami and geodetic data set (Fig. 3) provided


36 good control on the slip distribution (Supplementary Fig. S2). Both
amplitudes and timing of tsunami data are well predicted by the
1835

retrieved model, InSAR residuals are within 10% over most of the
covered area (see also Supplementary Fig. S3) and GPS vectors are
well reproduced, particularly in the epicentral region. Land-level
change data are predicted within errors almost everywhere as well
as the hinge-line for the coseismic displacement (Fig. 3b,c; see also
38 Fig. 2), which should run approximately along the coastline15 . Two
1960 other published models share first-order features with ours16,17 ,
and a third shares the main slip concentration north of the
epicentre18 . A comparison between their resolution analyses16,17 and
ours (Supplementary Fig. S2) shows that by including tsunami and
further geodetic data we achieve better resolution. We moreover
0 100 use variable strike and dip over the fault, which better captures
(km)
the geometrical complexity of the plate interface19 . A detailed
40
76 74 72 70 comparison with all of the published and preliminary online slip
Longitude (° W) models is reported in Supplementary Information.
In Fig. 4a we compare our slip model with past earthquakes
and the preseismic locking distribution13 . The gap zone roughly
0 5 10 15 20 corresponds to the 1835 source zone, where the locking is very high
Slip (m) (locally >85%), and no significant earthquakes have occurred since
then. The rupture was bilateral and propagated north and south
Figure 2 | Slip distribution of the 2010 Maule earthquake. Slip distribution from the hypocentre at an average velocity of 2–2.6 km s−1 (refs 16,
for the 2010 Mw 8.8 Maule earthquake obtained from the joint inversion of 18; fixed at 2.25 km s−1 in our model). The Maule earthquake
tsunami and geodetic data, represented by colours according to the scale at nucleated where locking is >75%, and propagated principally to
the bottom. White arrows represent the slip direction (rake). Thin black the north, releasing the maximum slip where the 1928 event had
contours indicate the associated surface vertical displacement partially relaxed the fault, and where the locking was lower around
(1-m-interval solid lines for uplift, 20-cm-interval dashed lines for 35◦ S. The rupture then propagated into a zone of higher locking
subsidence). Epicentres and source zones are plotted only for major thrust farther north and stopped near the southern termination of the 1985
earthquakes (compare Fig. 1). event, which could be expected to have low prestress.
The southward-propagating rupture, conversely, initiated with
(GPS) and land-level changes15 ) enable the estimation of the lower slip in a relatively weakly locked zone (50%–75%) before
onshore displacement field. The two data sets possess almost releasing more slip in a strongly locked area (>75%) around 37◦ S,
complementary resolving power on the coseismic deformation, and overlapping the 1960 source zone. The northern section of the
from which the characteristics of the earthquake rupture can be 1960 source region is known to have experienced relatively small
inferred. Here we carry out a joint inversion of these data to derive a slip and the termination of the Maule rupture coincides reasonably
robust model for the slip distribution, and the subsequent Coulomb well with the northern 1960 asperity at about 38.5◦ S (ref. 19).
stress change over the fault plane. The slip distribution of the Maule earthquake is then not
It is usually assumed that high prestress coincides with maxima consistent with simple earthquake recurrence models and our
of locking and elapsed time since the most recent failure. It is results contradict the recent work, on the basis of preliminary
also sometimes proposed that the slip distribution of a future slip distributions, that the slip distribution and preseismic locking
earthquake might be correlated with the distribution of prestress13 . are strongly correlated for this event13 . These preliminary slip
We then compare coseismic slip and stress change during the Maule models, however, were based on teleseismic data that, if used alone,
earthquake with the regional prestress distribution as inferred by seem to have poor resolution on the slip-distribution details16 .
preseismic locking and slip in past earthquakes. Our purpose is Further studies, including joint inversions of seismological, tsunami
twofold. We assess (1) the extent to which the slip distribution and geodetic data, might be necessary for better understanding
matches the inferred prestress distribution, to test the fundamental differences between our and published seismological models18 .
hypothesis that the two might be correlated, and (2) the extent to Similarly complex relationships between locking and slip can
which the Maule event closed the Darwin gap, reducing the seismic be nevertheless observed in other, well instrumented, recent
hazard in the region. subduction earthquakes20 . It is then not sufficient to consider
The slip distribution model obtained with the joint inversion previous earthquakes, plate interface coupling and approximately
of tsunami and geodetic data is shown in Fig. 2 (slip uncertainties uniform and constant loading with strain accumulated in the

2 NATURE GEOSCIENCE | ADVANCE ONLINE PUBLICATION | www.nature.com/naturegeoscience


© 2011 Macmillan Publishers Limited. All rights reserved.
NATURE GEOSCIENCE DOI: 10.1038/NGEO1073 LETTERS
a 1.6 0.4 b 32
ANC ANT Obs Pred
20 cm
0.9 0.5 1m
ARI CLD
0.9 1.3
CAL COQ 34
1.9 0.3
COR
D32412
0.3 0.6

Latitude (° S)
IQU SFE
1.7 36
2.0
TAL VAL
1.3 0.3
WAI RIK
0.9 0.2
OWE EAS 38
0.3 0.04
D51406 D51426
0.05 Observed
D54401 Predicted 0 100
(km)
40
60 min 74 72 70
Longitude (° W)

c d Residuals
32
Observed

Ascending
Predicted
34
100

36 90

80
38
70

Relative error (%)


0 100
Latitude (° S)

(km) 60
40
50
32
40
Descending

34 30

20
36
10

0
38

0 100
(km)
40
¬1 0 1 2 3 74 72 70
M Longitude (° W)

Figure 3 | Comparison between observed and predicted data sets. a, Observed and predicted tsunami time series. Peak value (m) is indicated for each
station. Abbreviated station names are as in Supplementary Table S2. b, Observed and predicted GPS vectors. Contour lines of predicted vertical
displacement as in Fig. 2. Yellow squares indicate positions of land-level-change measurements. c, Observed and predicted land-level changes plotted
against latitude scale of b. Error bars for observed data are experimental uncertainties15 , whereas for predicted data they are calculated adding ±1σ errors
(Supplementary Table S1) to the average slip model. d, Residuals between observed and predicted InSAR LOS displacement expressed as percentages of
the observed data values.

interseismic period on strongly coupled asperities and released itself provides a poor forecast of the slip and in this case would
during the earthquake, generating the regions of highest slip. The have even failed to forecast the location experiencing the highest
lobe of strong coupling, which extends from 38.5◦ S to about 36◦ S, slip. These results suggest that strong coupling is a necessary
remained largely unbroken despite its location in the centre of but not sufficient condition for high slip in a single event. Slip
the Darwin gap, whereas the area that experienced the maximum must additionally be controlled by other factors such as friction
slip of nearly 20 m in the 2010 event overlaps the source region and fault rheology, structural heterogeneity, and the history of
of the 1928 earthquake, although it could only have accumulated slip on the fault, including aseismic slip, probably over the past
some 5–6 m of slip since then. It is true that most of the several seismic cycles.
regions of high slip in Fig. 4a correspond to regions of relatively The Coulomb stress distribution induced on the fault plane
high (>50%) coupling, which probably controls the long-term by this model of slip during the Maule earthquake is shown in
distribution of seismic moment release; however, coupling by Fig. 4b. The increase in Coulomb stress deeper than 45 km or

NATURE GEOSCIENCE | ADVANCE ONLINE PUBLICATION | www.nature.com/naturegeoscience 3


© 2011 Macmillan Publishers Limited. All rights reserved.
LETTERS NATURE GEOSCIENCE DOI: 10.1038/NGEO1073

a 32 b 32

1985 1985

34 34

1928 1928

Latitude (° S)

Latitude (° S)
36 36

5
1835

183
38 38
1960 1960

0 100 0 100
(km) (km)
40 40
76 74 72 70 76 74 72 70

Longitude (° W) Longitude (° W)

0 5 10 15 20 ¬5 ¬4 ¬3 ¬2 ¬1 0 1 2 3
Coulomb stress change (MPa)
Slip (m)

Figure 4 | Comparison of the Maule earthquake slip distribution and coseismic stress variation to preseismic locking and past earthquakes. a, Slip
distribution for the Maule earthquake compared with the estimated position of the Darwin gap, in the segment where the 1835 earthquake probably
occurred. Source zones of past thrust earthquakes as in Fig. 1. White lines are preseismic locking contours as in Fig. 1. The Darwin gap was only partially
filled and a zone of high preseismic locking remains unbroken. b, Coulomb stress changes associated with the Maule earthquake, resolved on the
mega-thrust surface. Positive stress changes may favour a future rupture. An increase of stress occurring after the Maule earthquake in the Darwin gap
might have increased the probability of a future earthquake there.

shallower than about 15 km in Fig. 4b is unlikely to have strong of Hawaii (http://ilikai.soest.hawaii.edu/uhslc/background.html) sea-level data
consequences for seismic and tsunami hazard. The deeper region centres. Data providers for each station are listed on these websites, and in
might be beyond the seismogenic zone limit17 , and large shallow Supplementary Table S2. Deep-ocean Assessment and Reporting of Tsunami
sensor data are from the National Oceanic and Atmospheric Administration data
slow earthquakes are unlikely at accretionary margins21 . It might centre (http://www.ndbc.noaa.gov/).
be expected that some afterslip has occurred in both regions but We use 25 stations’ data from continuous GPS sites managed in South
this has not been detected so far. Ongoing geodetic studies will America by the International Global Navigation Satellite Systems Service
help to address this issue. However, our model shows that not (http://igscb.jpl.nasa.gov/) and three stations’ data distributed by the French Centre
National de la Recherche Scientifique–Institut National des Sciences de l’Univers
only did the Darwin gap not rupture completely, but the Maule
(INSU, France) through the GPSCOPE portal (https://gpscope.dt.insu.cnrs.fr/),
earthquake also produced a strong increase in stress on the highly hosted at the Division Technique–INSU. Using GIPSY-OASIS II v.5.0
coupled lobe extending northward from about 37◦ S to 36◦ S. Such software23 , we applied a robust procedure24 to analyse GPS data, consisting
stress increases have been frequently shown to trigger further of a precise-point-positioning strategy followed by network ambiguity resolution,
large earthquakes22 . In this case, failure of the unbroken areas of and, then, by alignment of the daily solutions to ITRF 2005 (ref. 25). For each site,
static coseismic offsets are estimated by comparing the average position calculated
the gap, if they are highly coupled, might produce an event of in the 7–8 days before the earthquake with the position obtained by the daily
magnitude Mw 7.5–8. solution of the day of the earthquake. The far-field stations are used to define a
In summary, the Maule earthquake was not the characteristic stable reference frame. In the inversion, we use only the six GPS data closest to the
earthquake that was expected to close the Darwin gap. Whilst the source (Supplementary Table S3).
rupture does seem to terminate against the source regions of the InSAR enables imaging of the relative change of the ground deformation with
high resolution and wide coverage soon after the event. For repeat-pass
1960 and 1985 earthquakes and the high slip is broadly confined interferometry shorter time span between the satellite acquisitions is
to areas of relatively high coupling, zones of very high coupling preferred to isolate the coseismic signal from the preseismic and postseismic
in the Darwin gap remain unbroken. Moreover, the highest slip deformation. In the present case the acquisition date ranges from May 2007
occurs in an area that failed in 1928 that is coupled only at the to February 2010 before the earthquake, and March 2010 to May 2010
50–75% level. It is then hard to envisage that preseismic locking after the earthquake17 . We use ascending and descending L-band images
acquired by ALOS PALSAR, a satellite of Japan Aerospace Exploration Agency
might be used for anticipating future slip in seismic gaps. The area (http://www.jaxa.jp/index_e.html), and this provides two Line-Of-Sight (LOS)
of strong plate locking in the Darwin gap that did not fail in this components of the three-dimensional displacement field (Supplementary Fig. S3).
event experienced some 2 MPa of further loading by it. Rather than We calibrated InSAR LOS data by means of the GPS data set described above (see
relaxing accumulated stress in the Darwin gap, reducing near-future also http://supersites.unavco.org/chile.php and ftp://topex.ucsd.edu/pub/chile_eq).
seismic hazard there, this strong stress interaction might have We use 34 land-level-change data observed mostly at coastal sites and some
at estuarine valleys15 . These data are an estimate, at several sites, of the coseismic
increased the probability of another major to great earthquake in displacement produced by the 27 February 2010 earthquake, spanning almost the
the Darwin gap in the near future. whole source length (Fig. 3c).
Our fault model has variable dip and strike. We sample the
Methods cross-section of the subduction zone geometry analysis carried
We use sea-level recordings at 15 coastal tide-gauges around the Pacific out by Gavin Hayes at the United States Geological Survey
Ocean, and at four deep-sea bottom pressure gauges in the open sea (http://earthquake.usgs.gov/earthquakes/eqarchives/subduction_
(Supplementary Fig. S4). Tide-gauge data are from Intergovernmental zone/us2010tfan/), and then develop it along strike, using trench coordinates
Oceanographic Commission/United Nations Educational, Scientific and provided by Peter Bird26 . The fault plane is divided into 200 subfaults of 25×25 km
Cultural Organization (http://www.ioc-sealevelmonitoring.org/) and University (Supplementary Fig. S5 and Table S1).

4 NATURE GEOSCIENCE | ADVANCE ONLINE PUBLICATION | www.nature.com/naturegeoscience


© 2011 Macmillan Publishers Limited. All rights reserved.
NATURE GEOSCIENCE DOI: 10.1038/NGEO1073 LETTERS
The vertical displacement associated with each of the subfaults is 14. Darwin, C. Journal of the Researches into the Natural History and Geology of the
calculated with Okada’s formulas27 , and transferred to the sea surface Countries Visited During the Voyage of the HMS Beagle Round the World 2nd
as the initial condition for tsunami propagation. The propagation is edn (John Murray, 1845).
calculated with the Cornell Multi-Grid Coupled Tsunami Model code 15. Farías, M. et al. Land-level changes produced by the Mw 8.8 2010 Chilean
(http://ceeserver.cee.cornell.edu/pll-group/comcot.htm), based on shallow earthquake. Science 329, 916 (2010).
water equations in spherical coordinates. We use total reflecting boundaries 16. Delouis, B., Nocquet, J-M. & Vallée, M. Slip distribution of the February 27,
at the coastlines and open boundaries elsewhere. Linear equations are 2010 Mw = 8.8 Maule earthquake, central Chile, from static and high-rate GPS,
solved for the propagation in the open sea, using a grid spacing of InSAR, and broadband teleseismic data. Geophys. Res. Lett. 37, L17305 (2010).
2 arcmin. For coastal propagation we use nonlinear equations, and 17. Tong, X. et al. The 2010 Maule, Chile earthquake: Downdip rupture limit
finer nested grids with 30 arcsec spacing. We use the SRTM30_PLUS revealed by space geodesy. Geophys. Res. Lett. 37, L24311 (2010).
(http://topex.ucsd.edu/WWW_html/srtm30_plus.html) bathymetric model. 18. Lay, T. et al. Teleseismic inversion for rupture process of the 27 February 2010
Okada’s formulas are also used for calculating subfault contributions at GPS, Chile (Mw 8.8) earthquake. Geophys. Res. Lett. 37, L13301 (2010).
land-level data sites and InSAR, for a set of points subsampled from the ALOS 19. Moreno, M. S., Bolte, J., Klotz, J. & Melnick, D. Impact of megathrust geometry
PALSAR interferograms. A variable LOS based on the satellite state vectors is used on inversion of coseismic slip from geodetic data: Application to the 1960 Chile
for the projection into the ground-satellite direction. earthquake. Geophys. Res. Lett. 36, L16310 (2009).
The inversion method28 is based on a global search technique, simulated 20. Konca, O. et al. Partial rupture of a locked patch of the Sumatra megathrust
annealing and the linear superposition of Green’s functions. We add smoothing during the 2007 earthquake sequence. Nature 456, 631–635 (2008).
and seismic-moment-minimization constraints to the slip distribution. The 21. Bilek, S. L. The role of subduction erosion on seismicity. Geology 38,
rupture front is assumed to be circular and propagating at constant speed, fixed at 479–480 (2010).
2.25 km s−1 (ref. 18), as we had poor resolution on this parameter. The slip vector 22. Nalbant, S. S., Steacy, S., Sieh, K., Natawidjaja, D. & McCloskey, J. Earthquake
(slip, rake) for each subfault is the observable retrieved with the inversion (Fig. 2). risk on the Sunda trench. Nature 435, 756–757 (2005).
The rake is kept constant over three large blocks (whose centres are the starting 23. Lichten, S. & Borders, J. Strategies for high-precision Global Positioning
points of the white arrows in Fig. 2). We also make checkerboard tests for assessing System orbit determination. J. Geophys. Res. 92, 12751–12762 (1987).
the resolution of the single data sets and of both data sets together in the joint 24. D’Agostino, N. et al. Active tectonics of the adriatic region from GPS and
inversion (Supplementary Fig. S2). earthquake slip vectors. J. Geophys. Res. 113, B12413 (2008).
Coulomb stress values are computed by means of COULOMB 3.1 (refs 29,30). 25. Altamimi, Z., Collilieux, X., Legrand, J., Garayt, B. & Boucher, C. ITRF2005: A
Coulomb stress change is defined as 1CFF = 1τ +µ1σ , where 1τ is the change in new release of the international terrestrial reference frame based on time series
shear stress on the failure plane (positive in the slip direction), 1σ is the change in of station positions and earth orientation parameters. J. Geophys. Res. 112,
normal stress (positive when the fault is unclamped) and µ is the apparent friction B09401 (2007).
coefficient. The friction coefficient used in this study is = 0.4. 26. Bird, P. An updated digital model of plate boundaries.
Geochem. Geophys. Geosyst. 4, 1027 (2003).
Received 27 July 2010; accepted 31 December 2010; 27. Okada, Y. Surface deformation due to shear and tensile faults in a half-space.
published online 30 January 2011 Bull. Seismol. Soc. Am. 75, 1135–1154 (1985).
28. Lorito, S., Piatanesi, A., Cannelli, V., Romano, F. & Melini, D. Kinematics
References and source zone properties of the 2004 Sumatra–Andaman earthquake and
1. Barrientos, S. Is the Pichilemu–Talcahuano (Chile) a seismic gap? tsunami: Nonlinear joint inversion of tide gauge, satellite altimetry, and GPS
Seismol. Res. Lett. 61, 43 (1990). data. J. Geophys. Res. 115, B02304 (2010).
2. Campos, J. & Kausel, E. The large 1939 intraplate earthquake of Southern Chile. 29. Lin, J. & Stein, R. S. Stress triggering in thrust and subduction earthquakes,
Seismol. Res. Lett. 61, 43 (1990). and stress interaction between the southern San Andreas and nearby thrust
3. Madariaga, R. La Seismicidad de Chile, Fisica de la Tierra, Vol. 10 221–258 and strike-slip faults. J. Geophys. Res. 109, B02303 (2004).
(Ediciones de la Universidad Complutense de Madrid, 1998). 30. Toda, S., Stein, R. S., Richards-Dinger, K. & Bozkurt, S. Forecasting the
4. Beck, S., Barrientos, S., Kausel, E. & Reyes, M. Source characteristics of historic evolution of seismicity in southern California: Animations built on earthquake
earthquakes along the central Chile subduction zone. J. South Am. Earth Sci. stress transfer. J. Geophys. Res. 110, B05S16 (2005).
11, 115–129 (1998).
5. Klotz, J. et al. Earthquake cycle dominates contemporary crustal deformation Acknowledgements
in Central and Southern Andes. Earth Planet. Sci. Lett. 193, 437–446 (2001). We acknowledge discussions with our colleagues E. Tinti and A. Herrero about seismic
6. Ruegg, J. C. et al. Interseismic strain accumulation in south central Chile from rupture properties, and with S. Nalbant about coseismic stress. We also acknowledge
GPS measurements, 1996–1999. Geophys. Res. Lett. 29, 1517–1520 (2002). C. Vigny, leader for the acquisition of the GPSCOPE GPS data in Chile. We thank
7. Campos, J. et al. A seismological study of the 1835 seismic gap in South Central N. D’Agostino and E. D’Anastasio, who set up and implemented the GPS data processing
Chile. Phys. Earth Planet. Iner. 132, 177–195 (2002). strategy. We appreciate the effort of our colleagues at Cornell University who developed
8. Brooks, B. A. et al. Crustal motion in the Southern Andes (26◦ –36◦ S): the tsunami-modelling package. We moreover wish to thank all of the data providers
Do the Andes behave like a microplate? Geochem. Geophys. Geosyst. 4, who made this study possible. Some figures were drawn with generic mapping tools
1085 (2003). (http://gmt.soest.hawaii.edu/). J.M. acknowledges support from the UK NERC under
9. Moreno, M. S., Klotz, J., Melnick, D, Echtler, H. & Bataille, K. Active faulting grant numbers NE/F01161X/1 and NE/H008519/1.
and heterogeneous deformation across a megathrust segment boundary
from GPS data, south central Chile (36–39◦ S). Geochem. Geophys. Geosyst. 9,
Q12024 (2008).
Author contributions
10. Vigny, C. et al. Upper plate deformation measured by GPS in the Coquimbo S.L., F.R. and A.P. were involved in all of the phases of this study. S.A., X.T. and A.A.
Gap, Chile. Phys. Earth Planet. Iner. 175, 86–95 (2009). processed, modelled and analysed geodetic data, and wrote part of the Methods. J.M.
11. Ruegg, J. C. et al. Interseismic strain accumulation measured by GPS and M.C. contributed to result interpretation and paper writing. E.B. promoted the
in the seismic gap between Constitución and Concepción in Chile. experiment, contributed to result interpretation and supported the project.
Phys. Earth Planet. Iner. 175, 78–85 (2009).
12. Madariaga, R., Métois, M., Vigny, C. & Campos, J. Central chile finally breaks. Additional information
Science 328, 181–182 (2010). The authors declare no competing financial interests. Supplementary information
13. Moreno, M., Rosenau, M. & Oncken, O. 2010 Maule earthquake slip accompanies this paper on www.nature.com/naturegeoscience. Reprints and permissions
correlates with pre-seismic locking of Andean subduction zone. Nature 467, information is available online at http://npg.nature.com/reprintsandpermissions.
198–204 (2010). Correspondence and requests for materials should be addressed to S.L.

NATURE GEOSCIENCE | ADVANCE ONLINE PUBLICATION | www.nature.com/naturegeoscience 5


© 2011 Macmillan Publishers Limited. All rights reserved.

Вам также может понравиться