The April 1, 2014 Iquique, Chile M w 8.1 earthquake rupture sequence

Similar documents
Rupture Characteristics of Major and Great (M w 7.0) Megathrust Earthquakes from : 1. Source Parameter Scaling Relationships

Supplementary Material

SUPPLEMENTARY INFORMATION

Magnitude 8.2 NORTHWEST OF IQUIQUE, CHILE

Centroid moment-tensor analysis of the 2011 Tohoku earthquake. and its larger foreshocks and aftershocks

Originally published as:

Supplementary Materials for

JCR (2 ), JGR- (1 ) (4 ) 11, EPSL GRL BSSA

A GLOBAL SURGE OF GREAT EARTHQUAKES FROM AND IMPLICATIONS FOR CASCADIA. Thorne Lay, University of California Santa Cruz

Possible large near-trench slip during the 2011 M w 9.0 off the Pacific coast of Tohoku Earthquake

Centroid-moment-tensor analysis of the 2011 off the Pacific coast of Tohoku Earthquake and its larger foreshocks and aftershocks

Data Repository of Paper: The role of subducted sediments in plate interface dynamics as constrained by Andean forearc (paleo)topography

Rapid source characterization of the 2011 M w 9.0 off the Pacific coast of Tohoku Earthquake

Along-dip seismic radiation segmentation during the 2007 M w 8.0 Pisco, Peru earthquake

ABSTRACT. Key words: InSAR; GPS; northern Chile; subduction zone.

RELOCATION OF THE MACHAZE AND LACERDA EARTHQUAKES IN MOZAMBIQUE AND THE RUPTURE PROCESS OF THE 2006 Mw7.0 MACHAZE EARTHQUAKE

Slip distributions of the 1944 Tonankai and 1946 Nankai earthquakes including the horizontal movement effect on tsunami generation

Source rupture process of the 2003 Tokachi-oki earthquake determined by joint inversion of teleseismic body wave and strong ground motion data

Magnitude 7.1 NEAR THE EAST COAST OF HONSHU, JAPAN

Source Characteristics of Large Outer Rise Earthquakes in the Pacific Plate

SOURCE MODELING OF RECENT LARGE INLAND CRUSTAL EARTHQUAKES IN JAPAN AND SOURCE CHARACTERIZATION FOR STRONG MOTION PREDICTION

Seismic-afterslip characterization of the 2010 M W 8.8 Maule, Chile, earthquake based on moment tensor inversion

by Bertrand Delouis, Mario Pardo, Denis Legrand, * and Tony Monfret Introduction

The 2011 M w 9.0 off the Pacific coast of Tohoku Earthquake: Comparison of deep-water tsunami signals with finite-fault rupture model predictions

Geodesy (InSAR, GPS, Gravity) and Big Earthquakes

Technical Report December 25, 2016, Mw=7.6, Chiloé Earthquake

Study megathrust creep to understand megathrust earthquakes

SOURCE INVERSION AND INUNDATION MODELING TECHNOLOGIES FOR TSUNAMI HAZARD ASSESSMENT, CASE STUDY: 2001 PERU TSUNAMI

PUBLICATIONS. Geophysical Research Letters

Preliminary slip model of M9 Tohoku earthquake from strongmotion stations in Japan - an extreme application of ISOLA code.

EARTHQUAKE SOURCE PARAMETERS FOR SUBDUCTION ZONE EVENTS CAUSING TSUNAMIS IN AND AROUND THE PHILIPPINES

Empirical Green s Function Analysis of the Wells, Nevada, Earthquake Source

MMA Memo No National Radio Astronomy Observatory. Seismicity and Seismic Hazard at MMA site, Antofagasta, Chile SERGIO E.

Case Study 2: 2014 Iquique Sequence

Rapid Seismological Quantification of Source Parameters of the 25 April 2015 Nepal Earthquake

Triggering of earthquakes during the 2000 Papua New Guinea earthquake sequence

Widespread Ground Motion Distribution Caused by Rupture Directivity during the 2015 Gorkha, Nepal Earthquake

Tsunami waveform analyses of the 2006 underthrust and 2007 outer-rise Kurile earthquakes

Depth-varying rupture properties of subduction zone megathrust faults

27th Seismic Research Review: Ground-Based Nuclear Explosion Monitoring Technologies

overlie the seismogenic zone offshore Costa Rica, making the margin particularly well suited for combined land and ocean geophysical studies (Figure

Coseismic slip distribution of the 1946 Nankai earthquake and aseismic slips caused by the earthquake

Rapid magnitude determination from peak amplitudes at local stations

TSUNAMI HAZARD ASSESSMENT FOR THE CENTRAL COAST OF PERU USING NUMERICAL SIMULATIONS FOR THE 1974, 1966 AND 1746 EARTHQUAKES

A search for seismic radiation from late slip for the December 26, 2004 Sumatra-Andaman (M w = 9.15) earthquake

Source of the July 2006 West Java tsunami estimated from tide gauge records

Tsunami and earthquake in Chile Part 2

Seismic Activity near the Sunda and Andaman Trenches in the Sumatra Subduction Zone

Imaging short-period seismic radiation from the 27 February 2010 Chile (M W 8.8) earthquake by back-projection of P, PP, and PKIKP waves

RELOCATION OF LARGE EARTHQUAKES ALONG THE PHILIPPINE FAULT ZONE AND THEIR FAULT PLANES

Magnitude 7.6 SOUTH OF IQUIQUE, CHILE

Synthetic sensitivity analysis of high frequency radiation of 2011 Tohoku-Oki (M W 9.0) earthquake

Moment tensor inversion of near source seismograms

Aftershock seismicity of the 2010 Maule Mw=8.8, Chile, earthquake: Correlation between co-seismic slip models and aftershock distribution?

Sendai Earthquake NE Japan March 11, Some explanatory slides Bob Stern, Dave Scholl, others updated March

The Kuril Islands Great Earthquake Sequence

ON NEAR-FIELD GROUND MOTIONS OF NORMAL AND REVERSE FAULTS FROM VIEWPOINT OF DYNAMIC RUPTURE MODEL

Systematic deficiency of aftershocks in areas of high coseismic slip for large subduction zone earthquakes

Coseismic slip distribution of the 2005 off Miyagi earthquake (M7.2) estimated by inversion of teleseismic and regional seismograms

Challenges of Applying Ground Motion Simulation to Earthquake Engineering

This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and

Characteristics of seismic activity before Chile M W 8.8 earthquake in 2010

Rupture Process of the Great 2004 Sumatra-Andaman Earthquake

Contents of this file

The Size and Duration of the Sumatra-Andaman Earthquake from Far-Field Static Offsets

Source Process and Constitutive Relations of the 2011 Tohoku Earthquake Inferred from Near-Field Strong-Motion Data

PUBLICATIONS. Geophysical Research Letters

Scaling relations of seismic moment, rupture area, average slip, and asperity size for M~9 subduction-zone earthquakes

TSUNAMI CHARACTERISTICS OF OUTER-RISE EARTHQUAKES ALONG THE PACIFIC COAST OF NICARAGUA - A CASE STUDY FOR THE 2016 NICARAGUA EVENT-

The 2010 Mw 8.8 Chile earthquake: Triggering on multiple segments and frequency dependent rupture behavior

Dynamic source inversion for physical parameters controlling the 2017 Lesvos earthquake

Outer trench-slope faulting and the 2011 M w 9.0 off the Pacific coast of Tohoku Earthquake

Rapid Earthquake Rupture Duration Estimates from Teleseismic Energy Rates, with

MECHANISM OF THE 2011 TOHOKU-OKI EARTHQUAKE: INSIGHT FROM SEISMIC TOMOGRAPHY

Routine Estimation of Earthquake Source Complexity: the 18 October 1992 Colombian Earthquake

Earthquakes and Tsunamis

Originally published as:

Rupture process of the 2010 M w 7.8 Mentawai tsunami earthquake from joint inversion of near-field hr-gps and

Joint inversion of strong motion, teleseismic, geodetic, and tsunami datasets for the rupture process of the 2011 Tohoku earthquake

RELOCATION OF LARGE EARTHQUAKES ALONG THE SUMATRAN FAULT AND THEIR FAULT PLANES

California foreshock sequences suggest aseismic triggering process

The 2017 M W 8.2 Chiapas, Mexico Earthquake: Energetic Slab Detachment

Three Dimensional Simulations of Tsunami Generation and Propagation

revised October 30, 2001 Carlos Mendoza

Earthquake Source Dynamics and seismic radiation From large Chilean earthquakes

SUPPLEMENTARY INFORMATION

LETTER Earth Planets Space, 56, , 2004

SUPPLEMENTARY INFORMATION

Rupture complexity of the M w 8.3 sea of okhotsk earthquake: Rapid triggering of complementary earthquakes?

The Sanriku-Oki low-seismicity region on the northern margin of the great 2011 Tohoku-Oki earthquake rupture

Effect of an outer-rise earthquake on seismic cycle of large interplate earthquakes estimated from an instability model based on friction mechanics

AVERAGE AND VARIATION OF FOCAL MECHANISM AROUND TOHOKU SUBDUCTION ZONE

Title. Author(s)Heki, Kosuke. CitationScience, 332(6036): Issue Date Doc URL. Type. File Information. A Tale of Two Earthquakes

Distribution of slip from 11 M w > 6 earthquakes in the northern Chile subduction zone

Tsunami modeling from the seismic CMT solution considering the dispersive effect: a case of the 2013 Santa Cruz Islands tsunami

Effect of the Emperor seamounts on trans-oceanic propagation of the 2006 Kuril Island earthquake tsunami

NUMERICAL SIMULATIONS FOR TSUNAMI FORECASTING AT PADANG CITY USING OFFSHORE TSUNAMI SENSORS

Tsunami waves swept away houses and cars in northern Japan and pushed ships aground.

Magnitude 7.1 PERU. There are early reports of homes and roads collapsed leaving one dead and several dozen injured.

Megathrust Earthquakes

Transcription:

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 The April 1, 2014 Iquique, Chile M w 8.1 earthquake rupture sequence Thorne Lay 1,*, Han Yue 1, Emily E. Brodsky 1, and Chao An 2 1 Department of Earth and Planetary Sciences, University of California Santa Cruz, Santa Cruz, CA 95064, USA. 2 School of Civil and Environmental Engineering, Cornell University, Ithaca, NY 14853, USA *Corresponding author: Thorne Lay (tlay@ucsc.edu; 831-459-3164) Abstract. On April 1, 2014, a great (M w 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 backprojections 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 10 21 Nm with a fault dip of 18, radiated seismic energy of 4.5-8.4 10 16 J, and static stress drop of ~2.5 MPa. Most of the 1877 gap remains unbroken and hazardous. 1. Introduction Northern Chile experienced a great subduction zone megathrust earthquake on May 9, 1877 with estimated seismic magnitude of 8.7-8.9 [Comte and Pardo, 1991] and a tsunami magnitude M t of 9.0. Recent geodetic measurements of eastward deformation of the upper plate indicate that most of the 1877 rupture zone (Figure 1) from 19 S to 23 S has a high coupling coefficient, albeit with some patchiness along-strike and along-dip [e.g., Béjar-Pizarro et al., 2013; Métois et al., 2013]. This region has been identified as the north Chile seismic gap [Kelleher, 1972; Nishenko, 1985] based on the lack of large earthquakes for the 137 years over which the Nazca plate has been underthrusting South America at about 65 mm/yr [DeMets et al., 2010]. On the 1

26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 order of 6 to 9 m of slip deficit may have accumulated since 1877. The earthquake history prior to 1877 is uncertain [Nishenko, 1985; Comte and Pardo, 1991], so it is unclear whether the region regularly fails in huge single ruptures or intermittently in larger ruptures then sequences of smaller ruptures, as is the case along the Ecuador-Colombia coastline [Kanamori and McNally, 1982]. The rupture zone of the 1868 Peru earthquake, with an estimated seismic magnitude of 8.5-8.8, partly re-ruptured in the June 23, 2011 M w 8.4 Peru earthquake (Figure 1), leaving an ~100 km long region offshore of southeastern Peru just north of the 1877 gap that may also have large slip deficit [e.g., Loveless et al., 2010]. On March 16, 2014, an M w 6.7 thrust event occurred on or near the megathrust about 60 km north-northwest of Iquique, Chile, and was followed by two weeks of thrust aftershocks that slowly migrated (~20 km/week) northward along the megathrust from 20.2 S to 19.6 S. The location of this sequence in the northern portion of the 1877 seismic gap focused attention on the region, and on April 1, 2014 an M w 8.1 interplate thrust earthquake initiated at the northern end of the foreshock sequence (19.642 S, 70.817 W, 23:46:46 UTC [USGS National Earthquake Information Center NEIC: http://earthquake.usgs.gov/regional/neic/]). The global Centroid- Moment Tensor (gcmt) solution for this event [http://www.globalcmt.org/cmtsearch.html] indicates an almost purely double-couple faulting geometry with strike 357, dip 18, and rake 109, at a centroid depth of 21.9 km and centroid location south of the hypocenter (19.77 S, 70.98 W), with a centroid time shift of 42.5 s and seismic moment of 1.69 10 21 Nm (M w 8.1) (Figure 1). A substantial aftershock sequence ensued, the largest of which occurred on April 3, 2014 with M w 7.7 (02:43:14 UTC, 20.518 S, 70.498 W, centroid depth 31.3 km), 49 km southwest of Iquique. The region extending from up-dip of the April 3 event southward to ~23 S, up-dip of the 2007 M w 7.7 Tocopilla earthquake rupture zone [e.g., Béjar-Pizarro et al., 2

49 50 2010; Schurr et al., 2012; Contreras-Reyes et al., 2012], remains strained and has potential for either an ~M w 8.5 event or several smaller great events (Figure 2). 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 2. Quantifying the 2014 M w 8.1 Earthquake Sequence The north Chile megathrust region from 19 S to 21 S has more frequent large earthquakes in seismological catalogs such as PAGER-CAT [Allen et al., 2009] dating back to 1900 than adjacent portions of the 1868 and 1877 seismic gaps (Figure 2). Some of these events are likely intraplate based on their rather uncertain locations (Auxiliary Figure S1). Following the 2001 Peru earthquake, the rate of activity for events with magnitude 4.5 increased in the vicinity of the 2014 sequence. Focal mechanisms from the gcmt catalog dating back to 1976 are sparse along the southern portion of the 1868 rupture zone and the 1877 zone other than for the 2014 thrusting sequence (Figure S1). A W-phase inversion was performed for the 2014 mainshock using 160 ground motion recordings from 112 global seismic stations with stable signals in the 200 to 1000 s period range. This yielded an almost purely double-couple solution with best double-couple strike 348.7, dip 14, and rake 90.7, with moment of 1.99 10 21 Nm (M w 8.1), centroid time shift of 45 s, depth of 30.5 km and location 20.242 S, 70.710 W (Figure 1). This is compatible with the gcmt solution. Teleseismic P-waves from the 2014 mainshock recorded at uniformly-spaced North American broadband stations, were aligned by multi-station cross-correlation coefficient of the early P wave energy, filtered in the period range 2.0-0.5 s, and then back-projected to the source region following the procedure of Xu et al. [2009]. The normalized time-integrated beam power for integrations over 10 s moving windows with time shifts of 1 s is shown in Figure 3 (an 3

71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 animation of the time sequence is shown in a Supplemental Movie). The imaged short-period energy release region is surprisingly concentrated, extending along strike about 100 km, and located below the coastline down-dip from the hypocenter. The foreshock and aftershock sequences locate up-dip of this short-period source region. Teleseismic body wave inversions for a large rupture with concentrated slip tend to have strong trade-offs between rupture velocity and spatial extent of slip, due to lack of resolved directivity. We sought to define a priori bounds on the permissible rupture extent using tsunami observations from deep-water pressure sensors of the NOAA and Chilean DART systems. Three high-quality recordings of the mainshock tsunami (Figure 4) were used. A fault model based on the megathrust morphology inferred from the trench curvature and regional gravity modeling [Tassara and Echaurren, 2012] that extended to the trench and along strike several degrees (with ~40 km spacing along strike and 20-40 km spacing along dip), was used in an initial inversion. The global seismic data set is comprised of 52 P waves and 49 SH wave broadband ground displacements in the period range 1.1 to 200 s. The same data are used in inversions of only seismic waves and jointly with the tsunami observations. Subfault rupture durations of 40 s were used for the large grid spacing, which was designed to be compatible with the tsunami wavelength sensitivity. A rupture velocity of 2.2 km/s was assumed. The seismic wave-only inversion, using typical regularization, tends to distribute moment spatially proportional to rupture velocity for the kinematic inversion, even given the long subfault durations, resulting in distributed slip across the fault plane, including extending to near the trench (Figure 4, Model A). Green functions for the tsunami signals are computed for the model grid using the linear version of the Cornell Multi-grid Coupled Tsunami model code COMCOT, with additional corrections for waveguide and elastic dispersion and water density variations [Tsai et al., 2013] 4

94 95 96 97 98 99 100 101 102 103 104 105 106 107 108 109 110 111 112 113 114 115 116 being applied. The bathymetry model had a spatial resolution of 2 arcmin, which was resampled from the GEBCO_08 data with a resolution of 30 arcsec. Inclusion of the three tsunami recordings, which are given relatively high weight in the joint inversion (a similar model is found by inverting only the tsunami data), suppresses the isochronal expansion of the seismic moment distribution, concentrating rupture in the subfaults near the hypocenter, and allowing inversion without regularization. Very shallow slip near the trench is not consistent with the timing or waveforms of the tsunami and overall slip is spatially concentrated down-dip of the hypocenter. Guided by the back-projections and tsunami inversion, we constructed a final preferred finite-fault model from linear least-squares inversion of only the seismic data [Kikuchi and Kanamori, 1992]. The rupture model assumes the gcmt strike (357 ) and dip (18 ) for a planar fault 157.5 km long and 105 km wide with 17.5 km grid spacing (Figure S2). The subfault source durations are parameterized with 6 2.5-s rise-time triangles offset by 2.5 s, giving possible subfault durations of 17.5 s. Rake was allowed to vary for each subfault subevent. Our final seismological model is shown in map view in Figure 1, with a detailed display in Figure S2. Waveform matches to the P and SH data account for about 88% of the signal power and are shown in Figure S3. The seismic moment is 1.66 x 10 21 Nm (M w 8.08), very close to the gcmt moment. The primary slip zone is about 30 km down-dip and 30 km southward along strike from the hypocenter, with slip of up to ~6.7 m (peak slip is not a well-resolved parameter) for this model parameterization (Figure 1). The average slip for subfaults with moment at least 15% of the peak subfault moment is 2.6 m, and the static stress drop for a circular fault model with corresponding area is 2.5 MPa. While there is some smearing of the moment distribution along circular isochrons, the imposition of a priori spatial constraints on the inversion leads to very stable results. The centroid time is 41.7 s. There is an ~15 s interval of weak teleseismic P wave 5

117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134 135 136 137 138 139 motion at the onset of this earthquake during which it is unclear whether the rupture expanded or concentrated at the edge of the large slip patch that ruptured subsequently. Given the imposed kinematic constraints, this model provides a seismological model for the earthquake consistent with very broadband teleseismic observations. This model predicts ENE direction of ground motion at GPS station IQQE (located in Figure 1) well, but over-predicts the offset by 60%. A finite-fault model based on teleseismic P and SH wave broadband ground displacements was also developed for the M w 7.7 aftershock on April 3, 2014. This rupture also has a very concentrated slip distribution. The same strike and dip as for the mainshock (which is the same as the gcmt solution for the April 3 event) are used with a hypocentral depth of 35 km, and a 9 9 grid with 10 km spacing. A rupture velocity of 1.75 km/s and moderately long subfault rupture durations comprised of 7 2-s rise-time triangles (16 s) were assumed to allow the rupture to concentrate spatially as much as indicated by the data. The resulting slip model (Figure S4) and waveform fits (Figure S5) account for 80% of the waveform power with a seismic moment of 4.27 10 20 Nm (M w 7.7) and a centroid time of 21 s. The slip area is located south of the mainshock slip zone, centered near the hypocenter (Figure 1). Our 200 to 600 s period W-phase inversion (based on 114 waveforms from 82 stations) for this event has a moment of 4.47 10 20 Nm, strike 356.6, dip 16.1, rake 100.3 and centroid time shift of 22 s and depth of 35.5 km. Radiated energy estimates and average source spectra were computed for the April 1 mainshock and April 3 aftershock. The total radiated energy for the mainshock is estimated to be 8.4 10 16 J, with a seismic moment-scaled energy ratio of 4.95 10-5. These estimates are based on average teleseismic P wave spectra for the frequency range 0.05 to 2.0 Hz for stations with radiation factors larger than 0.5, along with the energy contribution from lower frequencies based on the moment rate spectra for each event (Figure S6). If we restrict the data to those with 6

140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161 162 radiation factors larger than 0.7, the energy estimate reduces to 6.4 10 16 J, and if we further restrict to a maximum frequency of 1.0 Hz it is 4.5 10 16 J. The source spectrum is generally consistent with a standard interplate frequency-squared reference spectrum with a stress drop parameter of about 3 MPa (Figure S6). There is no clear indication in the averaged spectra of any slow slip component associated with the 15-s interval of weak seismic radiation. The April 3 aftershock has a radiated energy of 1.0 10 16 J, and the seismic moment-scaled energy ratio is 2.28 10-5. 3. Discussion and Conclusion The April 1, 2014 M w 8.1 Iquique earthquake ruptured about 20% of the seismic gap along the larger 1877 rupture. The remaining unbroken zone has now been bracketed by M w 8.1 events to the north and south (the April 1, 2014 Iquique and July 30, 1995 Antofagasta [e.g., Pritchard et al., 2002] events in Figure 2) and by M w 7.7 events somewhat down-dip on the northern and southern edges of the gap (the April 4, 2014 Iquique and November 14, 2007 Tocopilla events). Geodetic inversions for recent slip deficit suggest relatively uniform coupling in the unbroken region, so failure in a single event is certainly plausible, but the possibility of several smaller events releasing the strain cannot be ruled out. There is little sediment or bathymetric structure on the subducting plate along the remaining gap [Loveless et al., 2010]. The 2014 foreshock sequence is compared in Figure 5 with the foreshock sequence for the great 2011 Tohoku earthquake and a comparable size swarm in 1997 along central Chile that did not lead to a great earthquake. Is it viable to recognize a foreshock sequence like that in 2014 as a definite precursor to an imminent large event or not? A significant number of large interplate events appear to have had foreshock activity and possibly slow slip events over variable time periods before the large earthquakes [e.g., Bouchon et al., 2013], and it has been suggested that 7

163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183 184 185 this could assist in earthquake mitigation. Statistical attributes of the foreshock sequences may reveal anomalous rate increases if prior seismicity is well cataloged. The 2014 foreshock sequence does have unusually high rate increase and has several relatively large events after the magnitude 6.7 event that enhance the regional rate change. The 2011 Tohoku foreshock sequence does not have such a distinctive rate change in teleseismic catalogs, but detailed studies of local recordings reveal an immense number of events in the two days between the March 9, 2011 M w 7.3 foreshock and the mainshock [Kato et al., 2010; Marsan and Enescu, 2012]. Migrations of seismicity toward the mainshock epicenter at a few kilometers/day occurred in the month preceding the March 11 event and in the 15 to 18 hours after the 7.3 foreshock (Figure 5), which has some parallels to the observed northward migration of aftershocks toward the 2014 Iquique mainshock hypocenter in the preceding two weeks. These migrations tend to support the notion of some slow slip process concentrating stress near the mainshock hypocenter within weeks of the mainshock. The 2014 event differs from 2011 in not having sufficient down-dip slip to drive rupture up-dip to the trench. It is not clear if the shallow megathrust in northern Chile is seismogenic or not, and landbased slip-deficit estimates from geodesy lack resolution near the trench. Should the remaining 1877 gap rupture in a single event, it may drive slip shallower than the April 1, 2014 event. The very large tsunami in 1877 may be indicative of that type of broadening of the slip zone, as occurred in the 2011 Tohoku and 1906 Ecuador-Colombia events. The July 1997 earthquake sequence in central Chile, near the Coquimbo seismic gap, has a comparable largest event (6.7) to the Iquique foreshocks, and some spatial migration [Gardi et al., 2006; Holtkamp et al., 2011], but was followed by the October 15, 1997 M w 7.1 Punitaqui intraslab earthquake (normal faulting ~70 km deep) rather than a great interplate rupture [Pardo 8

186 187 188 189 190 191 192 193 et al., 2002]. This megathrust last failed in an M S 7.9 event in 1943, and prior to that possibly in 1880 [Beck et al., 1998], suggesting that a relatively large strain accumulation is possible. However, geodetic estimates of recent slip deficit in the Coquimbo gap region [Vigny et al., 2009; Métois et al., 2012] are low, so the Central Chile region may not be as close to its failure limit, leading to a swarm in a locally modestly coupled region that does not precipitate a larger failure. For the 2014 Iquique events Béjar-Pizaaro et al. [2013] and Métois et al. [2013] indicate stronger coupling on the deeper portion of the megathrust and less coupling up-dip near the Iquique events, consistent with the slowly migrating up-dip foreshock sequence followed by the 194 deeper down-dip mainshock. Holtkamp and Brudzinski [2011] examined many megathrust 195 196 197 198 199 200 201 202 earthquake swarms around Pacific subduction zones, finding that they tend to concentrate in regions of low interplate coupling or on the margins of large slip zones. Improved on-shore and off-shore geodetic and seismic networks in many regions are needed to establish the role of slow-slip events in foreshock sequences and better confidence in advance identification of likely foreshock sequences prior to great interplate events. Acknowledgments. The IRIS DMS data center was used to access the seismic data from Global Seismic Network and Federation of Digital Seismic Network stations. This work was supported by NSF grant EAR0635570 (T. L.). 203 9

204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 References Allen, T. I., K. D. Marano, P. S. Earle, and D. J. Wald (2009), PAGER-CAT: A composite earthquake catalog for calibrating global fatality models, Seism. Res. Lett., 80, 57-62, doi:10.1785/gssrl.80.1.57. Beck, S., S. Barrientos, E. Kausel, and M. Reyes (1998), Source characteristics of historic earthquakes along the central Chile subduction zone, J. South American Earth Sci., 11, 115-129. Béjar-Pizarro, M., et al. (2010), Asperities and barriers on the seismogenic zone in North Chile: state-of-the-art after the 2007 M w 7.7 Tocopilla earthquake inferred by GPS and InSAR data, Geophys. J. Int., 183, 390-406. Béjar-Pizarro, M., A. Socquet, R. Armijo, D. Carrizo, J. Genrich, and M. Simons (2013), Andean structural control on interseismic coupling in the North Chile subduction zone, Nature Geoscience, 6, 462-467, doi: 10.1038/NGEO1802. Bouchon, M., V. Durand, D. Marsan, H. Karabulut, and J. Schmittbuhl (2013), The long precursory phase of most large interplate earthquakes, Nature Geoscience, 6, 299-302. Comte, D., and M. Pardo, M. (1991), Reappraisal of great historical earthquakes in the northern Chile and southern Peru seismic gaps, Natural Hazards, 4(1), 23 44, doi:10.1007/bf00126557. Contreras-Reyes, E., J. Jara, I. Grevemeyer, S. Ruiz, and D. Carrizo (2012), An abrupt change in the dip of the subducting plate beneath north Chile, Nature Geosci., 5, 342-345, doi: 10.1038/NGEO1447. DeMets, C., Gordon, R. G., Argus, D. F. (2010), Geologically current plate motions. Geophys. J. Int., 181: 1 80. doi: 10.1111/j.1365-246X.2009.04491.x. 10

227 228 229 230 231 232 233 234 235 236 237 238 239 240 241 242 243 244 245 246 247 248 249 Gardi, A., A. Lemoine, R. Madariaga, and J. Campos (2006), Modeling of stress transfer in the Coquimbo region of central Chile, J. Geophys. Res., 111, B04307, doi:10.1029/2004jb003440. Holtkamp, S. G., and M. R. Brudzinski (2011), Earthquake swarms in the circum-pacific subduction zones, Earth Planet. Sci. Lett., 305, 215-225. Holtkamp, S. G., M. E. Pritchard, and R. B. Lohman (2011), Earthquake swarms in South America, Geophys. J. Int., 187, 128-146, doi: 10.1111/j.1365-246X.2011.05137.x. Kanamori, H., and K. McNally (1982), Variable rupture mode of the subduction zone along the Ecuador Colombia coast, Bull. Seismol. Soc. Am., 72, 1241 1253. Kato, A., K. Obara, T. Igarashi, H. Tsuruoka, S. Nakagawa, and N. Hirata (2012), Propagation of slow slip leading up to the 2011 M w 9.0 Tohoku-Oki earthquake, Science, 335, 705-708. Kelleher, J.A. (1972), Rupture zones of large South American earthquakes and some predictions. J. Geophys. Res., 77(11), 2087 2103. Kikuchi, M., H. Kanamori (1991). Inversion of complex body waves III. Bull. Seismol. Soc. Am., 81(6), 2335-2350. Lay, T., C. J. Ammon, A. R. Hutko, and H. Kanamori (2010). Effects of kinematic constraints on teleseismic finite-source rupture inversions: Great Peruvian earthquakes of 23 June 2001 and 15 August 2007, Bull. Seism. Soc. Am., 100, 969-994, doi:10.1785/0120090274. Loveless, J. P., M. E. Pritchard, and N. Kukowski (2010), Testing mechanisms of seismic segmentation with slip distributions from recent earthquakes along the Andean margin. Tectonophysics, 495, 15-33. Marsan, D., and B. Enescu (2012), Modeling the foreshock sequence prior to the 2011, Mw9.0 Tohoku, Japan, earthquake, J. Geophys. Res., 117, B06316, doi:10.1029/2011jb009039. 11

250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 Métois, M., A. Socquet, and C. Vigny (2012), Interseismic coupling, segmentation and mechanical behavior of the central Chile subduction zone, J. Geophys. Res., 117, B03406, doi:10.1029/2011jb008736. Métois, M., A. Socquet, C. Vigny, D. Carrizon, S. Peyrat, A. Delorme, E. Maureira, M.-C. Valderas-Bermejo, and I. Ortega (2013) Revisiting the North Chile seismic gap segmentation using GPS-derived interseismic coupling, Geophys. J. Int., 194, 1283-1294, doi: 10.1093/gji/ggt183. Nishenko, S. P. (1985), Seismic potential for large and great interplate earthquakes along the Chilean and southern Peruvian margins of South America: a quantitative reappraisal, J. Geophys. Res., 90, 3589 3615. Pardo, M., D. Comte, T. Monfret, R. Boroschek, and M. Astroza (2002), The October 15, 1997 Punitaqui earthquake (M w =7.1): a destructive event within the subducting Nazca plate in central Chile, Tectonophysics, 345, 199-210. Pritchard, M. E., M. Simons, P. A. Rosen, S. Hensley, and F. H. Webb (2002), Co-seismic slip from the 1995 July 30 M w =8.1 Antofagasta, Chile earthquakes as constrained by InSAR and GPS observations, Geophys. J. Int., 150, 362-376, doi:10.1046/j.1365-246x.2002.01661.x. Schurr, B., G. Asch, M. Rosenau, R. Wang, O. Oncken, S. Barrientos, P. Salazar, and J.-P. Vilotte (2012), The 2007 M7.7 Tocopilla northern Chile earthquake sequence: Implications for along-strike and downdip rupture segmentation and megathrust frictional behavior, J. Geophys. Res., 117, B05305, doi:10.1029/2011jb009030. Tassara, A., and A. Echaurren (2012), Anatomy of the Andean subduction zone: three dimensional density model upgraded and compared against global scale models, Geophysical Journal International, 189(1), 161-168. 12

273 274 275 276 277 278 279 280 281 Tsai, V. C., J.-P. Ampuero, H. Kanamori, and D. J. Stevenson (2013), Estimating the effect of Earth elasticity and variable water density on tsunami speeds, Geophys. Res. Lett., 40, 492-496, doi:10.1002/grl.50147. Vigny, C., A. Rudloff, J.-C. Ruegg, R. Madariaga, J. Campos, and M. Alvarez (2009), Upper plate deformation measured by GPS in the Coquimbo gap, Chile, Phys. Earth Planet. Inter., 175(1 2), 86 95. Xu, Y., K. D. Koper, O. Sufri, L. Zhu, A. R. Hutko (2009), Rupture imaging of the Mw 7.9 12 May 2008 Wenchuan earthquake from back projection of teleseismic P waves, Geochem. Geophys. Geosyst., 10, Q04006, doi:10.1029/2008gc002335. 282 13

283 284 285 286 287 288 289 290 291 292 293 294 295 Figure 1. Map of the subduction zone along which the Nazca plate underthrusts South America in southern Peru and northern Chile. The USGS-NEIC epicenter of the April 1, 2014 M w 8.1 earthquake (red star) and epicenters of foreshocks from March 16, 2014 to April 1, 2014 (gray circles) and aftershocks from April 1, 2014 to April 8, 2014 (green circles), are shown scaled with symbol size proportional to seismic magnitude. The focal mechanisms are global Centroid- Moment Tensor (gcmt) best double-couple solutions for the mainshock and the April 3, 2014 M w 7.7 aftershock. The preferred co-seismic slip distribution based on inversion of teleseismic body waves is contoured in the rectangular plot. Estimated rupture zones of historic large earthquakes in 1868 and 1877 are shown by dashed ellipses. The rupture zone of the June 23, 2001 M w 8.4 Peru earthquake [Lay et al., 2010] is also shown. The yellow arrows indicate motion of the Nazca plate relative to a fixed South America plate based on model MORVEL [DeMets et al., 2010]. The W-phase solution centroid location is indicated by the blue star. The red triangle is the location of GPS station IQQE. 296 297 298 299 300 301 302 303 304 Figure 2. Shallow (depths 70 km) seismicity from 1900 to 2014 with magnitude 4.5 from the USGS-NEIC earthquake catalog with pre-1973 events from PAGER-CAT [Allen et al., 2009] located within 200 km from the trench along southern Peru and northern Chile are plotted versus latitude. The detection level of earthquakes varies with time and is much higher prior to 1973. Symbol size is scaled relative to magnitude and large events are identified on the right, relative to the estimated 1868 and 1877 rupture zones shown in Figure 1. The latitude extent of seismic gaps in the southern portion of the 1877 zone and the overlapping portions of the southern 1868 and northern 1877 zones are indicated by the arrows. 305 14

306 307 308 309 310 311 Figure 3. Image of coherent short-period seismic energy release from the April 1, 2014 M w 8.1 Iquique earthquake, obtained by back-projection of teleseismic P wave ground displacements in the period range 0.5 to 2.0 s from North American stations. The time-integrated beam power is shown in the map to display the overall distribution of energy release. The peak beam amplitudes as a function of time is shown at the top. A time-varying sequence of spatial images is provided in a Supplemental Movie. 312 313 314 315 316 317 318 319 320 321 322 323 Figure 4. A preliminary inversion of teleseismic P and SH waves and 3 tsunami sea-level observations conducted to establish spatial dimensions of the April 1, 2014 M w 8.1 rupture. A relatively coarse spatial grid was used, defining subfaults extending all the way to the trench with varying strike, dip, and depth to match the megathrust geometry. (a) Inversion result using only seismic waves results in a concentration of slip across the fault plane to the south from the hypocenter. (b) Inversion result using the same seismic data as in (a) jointly with three tsunami observations from the locations shown in (c), with waveforms in (d). The predicted tsunami waveforms for the models in (a) and (b) are compared with the data in (d), showing that the widely distributed slip some of which extends to the trench is not consistent with the timing, amplitude or waveshape of the tsunami signals, and that concentrated slip near the hypocenter as in model (b) matches the tsunami signals well. 324 325 326 327 Figure 5. Comparison of three earthquake sequences with magnitudes 4.5, involving large interplate thrust faulting in seismic gap regions. The top panel is for the July 7, M w 6.7 Punitaqui swarm in Central Chile along the 1943 earthquake (Coquimbo) seismic gap; the middle panel is 15

328 329 for the 2011 Tohoku M w 9.0 foreshock sequence; and the bottom panel is for the April 2014 Iquique, Chile sequence. 330 16

331 332 333 334 335 336 337 338 339 340 341 342 343 344 345 Figure 1. Map of the subduction zone along which the Nazca plate underthrusts South America in southern Peru and northern Chile. The USGS-NEIC epicenter of the April 1, 2014 M w 8.1 earthquake (red star) and epicenters of foreshocks from March 16, 2014 to April 1, 2014 (gray circles) and aftershocks from April 1, 2014 to April 8, 2014 (green circles), are shown scaled with symbol size proportional to seismic magnitude. The focal mechanisms are global Centroid- Moment Tensor (gcmt) best double-couple solutions for the mainshock and the April 3, 2014 M w 7.7 aftershock. The preferred co-seismic slip distribution based on inversion of teleseismic body waves is contoured in the rectangular plot. Estimated rupture zones of historic large earthquakes in 1868 and 1877 are shown by dashed ellipses. The rupture zone of the June 23, 2001 M w 8.4 Peru earthquake [Lay et al., 2010] is also shown. The yellow arrows indicate motion of the Nazca plate relative to a fixed South America plate based on model MORVEL [DeMets et al., 2010]. The W-phase solution centroid location is indicated by the blue star. The red triangle is the location of GPS station IQQE. 17

346 347 348 349 350 351 352 353 354 355 Figure 2. Shallow (depths 70 km) seismicity from 1900 to 2014 with magnitude 4.5 from the USGS-NEIC earthquake catalog with pre-1973 events from PAGER-CAT [Allen et al., 2009] located within 200 km from the trench along southern Peru and northern Chile are plotted versus latitude. The detection level of earthquakes varies with time and is much higher prior to 1973. Symbol size is scaled relative to magnitude and large events are identified on the right, relative to the estimated 1868 and 1877 rupture zones shown in Figure 1. The latitude extent of seismic gaps in the southern portion of the 1877 zone and the overlapping portions of the southern 1868 and northern 1877 zones are indicated by the arrows. 18

356 357 358 359 360 361 362 363 364 Figure 3. Image of coherent short-period seismic energy release from the April 1, 2014 M w 8.1 Iquique earthquake, obtained by back-projection of teleseismic P wave ground displacements in the period range 0.5 to 2.0 s from North American stations. The time-integrated beam power is shown in the map to display the overall distribution of energy release. The peak beam amplitudes as a function of time is shown at the top. A time-varying sequence of spatial images is provided in a Supplemental Movie. 19

365 366 367 368 369 370 371 372 373 374 375 376 377 378 Figure 4. A preliminary inversion of teleseismic P and SH waves and 3 tsunami sea-level observations conducted to establish spatial dimensions of the April 1, 2014 Mw 8.1 rupture. A relatively coarse spatial grid was used, defining subfaults extending all the way to the trench with varying strike, dip, and depth to match the megathrust geometry. (a) Inversion result using only seismic waves results in a concentration of slip across the fault plane to the south from the hypocenter. (b) Inversion result using the same seismic data as in (a) jointly with three tsunami observations from the locations shown in (c), with waveforms in (d). The predicted tsunami waveforms for the models in (a) and (b) are compared with the data in (d), showing that the widely distributed slip some of which extends to the trench is not consistent with the timing, amplitude or waveshape of the tsunami signals, and that concentrated slip near the hypocenter as in model (b) matches the tsunami signals well. 20

379 380 381 382 383 384 385 Figure 5. Comparison of three earthquake sequences with magnitudes 4.5, involving large interplate thrust faulting in seismic gap regions. The top panel is for the July 7, M w 6.7 Punitaqui swarm in Central Chile along the 1943 earthquake (Coquimbo) seismic gap; the middle panel is for the 2011 Tohoku M w 9.0 foreshock sequence; and the bottom panel is for the April 2014 Iquique, Chile sequence. 21

386 387 388 389 390 391 392 393 394 395 396 397 The April 1, 2014 Iquique, Chile M w 8.1 earthquake rupture sequence Thorne Lay 1,*, Han Yue 1, Emily E. Brodsky 1, and Chao An 2 1 Department of Earth and Planetary Sciences, University of California Santa Cruz, Santa Cruz, CA 95064, USA. 2 School of Civil and Environmental Engineering, Cornell University, Ithaca, NY 14853, USA *Corresponding author: Thorne Lay (tlay@ucsc.edu; 831-459-3164) Supplementary Materials Figures S1-S6 Animation S1 22

398 399 400 401 402 403 404 405 406 407 Figure S1. Map of the source regions of the 1868 Peru and 1877 north Chile megathrust ruptures (ellipses). All events in the USGS-NEIC catalog (including PAGER-CAT) with reported magnitude larger than 7.0 are shown with circles scale to magnitude and color-coded for source depth. Only events less that 70 km deep are included. Focal mechanisms for events less than 60 km deep in the gcmt catalog are shown at the NEIC epicentral locations scaled and color-coded in the same way. The motion of the Nazca plate relative to fixed South America for model MORVEL [DeMets et al., 2010] is indicated by the yellow arrows. 23

408 409 410 411 412 413 414 415 416 417 Figure S2. The preferred finite-source model for the April 1, 2014 M w 8.1 Iquique mainshock based on a kinematic linear least-squares inversion of 101 teleseismic broadband P and SH waves in the period band 1.1-200 s. Corresponding waveform fits are shown in Figure S5. The moment rate function is shown at the top and the slip distribution and subfault source time function durations are shown below, with strength of slip color-coded and indicated by the vectors. The purple circles are 10 s intervals of the expanding rupture front with a rupture velocity of 1.5 km/s. The red star indicates the position of the hypocenter, corresponding to the map display of this model in Figure 1. 24

418 419 420 421 422 423 Figure S3. Waveform fits (observed black, red calculated) for teleseismic P and SH waves for the 2014 Iquique mainshock using the model shown in Figure S2. The station azimuths and epicentral distances are indicated below the station names, and the peak-to-peak ground motion in microns for each observation is labeled in blue. The data have normalized amplitudes with true relative amplitudes of predictions. 424 425 25

426 427 428 429 430 431 432 433 Figure S4. The preferred finite-source model for the April 3, 2014 M w 7.7 Iquique aftershock based on a kinematic linear least-squares inversion of 105 teleseismic broadband P and SH waves in the period band 1.1-200 s. Corresponding waveform fits are shown in Figure S5. The moment rate function is shown at the top and the slip distribution and subfault source time function durations are shown below, with strength of slip color-coded and indicated by the vectors. The purple circles are 10 s intervals of the expanding rupture front with a rupture velocity of 1.5 km/s. 434 26

435 436 437 438 439 440 Figure S5. Waveform fits (observed black, red calculated) for teleseismic P and SH waves for the 2014 Iquique aftershock using the model shown in Figure S4. The station azimuths and epicentral distances are indicated below the station names, and the peak-to-peak ground motion in microns for the data is labeled in blue. The data have normalized amplitudes with true relative amplitudes of predictions. 27

441 442 443 444 445 446 447 448 Figure S6. Average source spectra and normalized cumulative energy as a function of frequency for the April 1, 2014 Iquique mainshock (top row) and April 3, 2014 aftershock (lower row). The total radiated energy, Er, is computed for the broadband spectra by averaging individual estimates from broadband teleseismic P waves in the passband 0.05-2.0 Hz, and then scaling up for the proportion of energy at lower frequency. The source region parameters used in the calculation are shown for each case. 28