The Kuril Islands Great Earthquake Sequence

Similar documents
SUPPLEMENTARY INFORMATION

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

The Kuril Islands great earthquake sequence

Magnitude 8.3 SEA OF OKHOTSK

Rupture Process of the Great 2004 Sumatra-Andaman Earthquake

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

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

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

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

Supplementary Materials for

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

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

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

Magnitude 7.9 SE of KODIAK, ALASKA

Magnitude 8.2 NORTHWEST OF IQUIQUE, CHILE

Magnitude 7.1 NEAR THE EAST COAST OF HONSHU, JAPAN

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

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

Rapid Earthquake Rupture Duration Estimates from Teleseismic Energy Rates, with

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

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

Section Forces Within Earth. 8 th Grade Earth & Space Science - Class Notes

Earthquakes and Earthquake Hazards Earth - Chapter 11 Stan Hatfield Southwestern Illinois College

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

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

Source Characteristics of Large Outer Rise Earthquakes in the Pacific Plate

Seismological Aspects of the December 2004 Great Sumatra-Andaman Earthquake

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

Earthquake Stress Drops in Southern California

AVERAGE AND VARIATION OF FOCAL MECHANISM AROUND TOHOKU SUBDUCTION ZONE

Magnitude 7.5 PALU, INDONESIA

Section 19.1: Forces Within Earth Section 19.2: Seismic Waves and Earth s Interior Section 19.3: Measuring and Locating.

Three Dimensional Simulations of Tsunami Generation and Propagation

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

Teleseismic waveform modelling of the 2008 Leonidio event

EARTHQUAKE LOCATIONS INDICATE PLATE BOUNDARIES EARTHQUAKE MECHANISMS SHOW MOTION

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

Triggering of earthquakes during the 2000 Papua New Guinea earthquake sequence

Magnitude 7.6 & 7.6 PERU

Magnitude 7.0 N of ANCHORAGE, ALASKA

Lateral variation of the D 00 discontinuity beneath the Cocos Plate

Magnitude 7.6 & 7.4 SOLOMON ISLANDS

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

Analysis of Seismological and Tsunami Data from the 1993 Guam Earthquake

Earthquakes Chapter 19

I. Locations of Earthquakes. Announcements. Earthquakes Ch. 5. video Northridge, California earthquake, lecture on Chapter 5 Earthquakes!

Depth-varying rupture properties of subduction zone megathrust faults

7.1 FIJI 1, :57:22 UTC

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

2008 Monitoring Research Review: Ground-Based Nuclear Explosion Monitoring Technologies

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

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

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

Magnitude 7.0 PERU. This region of the Andes is a sparsely populated area, there were no immediate reports of injuries or damage.

Supporting Online Material for

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

revised October 30, 2001 Carlos Mendoza

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

Earthquakes. Building Earth s Surface, Part 2. Science 330 Summer What is an earthquake?

Magnitude 7.0 VANUATU

Imaging sharp lateral velocity gradients using scattered waves on dense arrays: faults and basin edges

Estimation of S-wave scattering coefficient in the mantle from envelope characteristics before and after the ScS arrival

SOURCE PROCESS OF THE 2003 PUERTO PLATA EARTHQUAKE USING TELESEISMIC DATA AND STRONG GROUND MOTION SIMULATION

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

Earthquakes Earth, 9th edition, Chapter 11 Key Concepts What is an earthquake? Earthquake focus and epicenter What is an earthquake?

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

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

Crustal deformation by the Southeast-off Kii Peninsula Earthquake

News Release December 30, 2004 The Science behind the Aceh Earthquake

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

1.3 Short Review: Preliminary results and observations of the December 2004 Great Sumatra Earthquake Kenji Hirata

Magnitude 6.5 OFFSHORE NORTHERN CALIFORNIA

Apparent Slow Oceanic Transform Earthquakes Due to Source Mechanism Bias

Magnitude 7.6 SOUTH OF IQUIQUE, CHILE

Magnitude 7.2 OAXACA, MEXICO

A GLOBAL MODEL FOR AFTERSHOCK BEHAVIOUR

Earthquake patterns in the Flinders Ranges - Temporary network , preliminary results

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

The Earthquake of Padang, Sumatra of 30 September 2009 scientific information and update

Global surface-wave tomography

Internal Layers of the Earth

Scaling of apparent stress from broadband radiated energy catalogue and seismic moment catalogue and its focal mechanism dependence

Magnitude 7.8 SCOTIA SEA

Magnitude 7.0 NEW CALEDONIA

Dynamic Triggering Semi-Volcanic Tremor in Japanese Volcanic Region by The 2016 Mw 7.0 Kumamoto Earthquake

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

Lessons from the 2004 Sumatra earthquake and the Asian tsunami

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

Dynamic Stress Drop of Recent Earthquakes: Variations within Subduction Zones

COULOMB STRESS CHANGES DUE TO RECENT ACEH EARTHQUAKES

Mechanics of Earthquakes and Faulting

Magnitude 7.7 QUEEN CHARLOTTE ISLANDS REGION

GNH7/GG09/GEOL4002 EARTHQUAKE SEISMOLOGY AND EARTHQUAKE HAZARD

High-Frequency Ground Motion Simulation Using a Source- and Site-Specific Empirical Green s Function Approach

Magnitude 7.1 PHILIPPINES

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

Magnitude 8.2 FIJI. A magnitude 8.2 earthquake occurred km (226.7 mi) E of Suva, Fiji at a depth of km (350 miles).

Short Note Source Mechanism and Rupture Directivity of the 18 May 2009 M W 4.6 Inglewood, California, Earthquake

Synthetic Seismicity Models of Multiple Interacting Faults

Structural heterogeneity in the megathrust zone and mechanism of the 2011 Tohoku oki earthquake (Mw 9.0)

Transcription:

The 2006-2007 Kuril Islands Great Earthquake Sequence Thorne Lay 1, Charles J. Ammon 2, Kevin Furlong 2, Alexander R. Hutko 1, Hiroo Kanamori 3, Luis Rivera 4, and Aaron A. Velasco 5 The southwestern half of the ~500-km-long seismic gap in the central Kuril Island arc subduction zone experienced two great earthquakes with extensive pre-shock and aftershock sequences in late 2006 to early 2007. The nature of seismic coupling in the gap had been uncertain due to the limited historical record of prior large events (one M S = 8.0 event in 1915) and the presence of distinctive upper plate and trench structure relative to adjacent regions along the arc that have experienced great interplate earthquakes in the last century. The outer rise region seaward of the seismic gap had experienced several compressional events over the preceding decades, leading to speculation that it was seismically coupled. This issue was partly resolved by failure of the shallow portion of the interplate megathrust in an M W = 8.3 thrust event on 15 November 2006. This event ruptured ~250 km along the seismic gap, northeast of the rupture zone of the 1963 Kuril Island earthquake. Within minutes of the thrust event, intense earthquake activity commenced in the outer rise seaward of the rupture zone, with the larger events having normal faulting mechanisms. An unusual double band of interplate and intraplate aftershocks developed. On 13 January 2007, an M W = 8.1 extensional earthquake ruptured within the Pacific plate beneath the seaward edge of the Kuril trench. This event is the third largest 1 Department of Earth and Planetary Sciences, University of California, Santa Cruz, California, USA 2 Department of Geosciences, The Pennsylvania State University, State College, Pennsylvania, USA 3 Seismological Laboratory, California Institute of Technology, Pasadena, California, USA 4 Institut de Physique du Globe de Strasbourg, France 5 Department of Geological Sciences, University of Texas, El Paso, Texas, USA - 1 -

normal-faulting outer rise earthquake on record, and its rupture zone extended tens of kilometers into the plate and parallel to most of the length of the 2006 thrust event rupture zone. It produced stronger shaking in Japan than the larger thrust event, as a consequence of greater short-period energy radiation from the source. The rupture process and slip distributions of the two great events are determined using very broadband teleseismic body and surface wave observations. The occurrence of the thrust event in the shallow portion of the interplate fault in a region with a paucity of large thrust events suggests that flexure of the upper plate primarily accommodated and released the strain accumulation. The inferred stress change on the plate is consistent with the stress drop involved in the outer rise normal faulting event, suggesting nearly complete stress drop on the interplate fault. This great earthquake doublet demonstrates the heightened seismic hazard posed by induced outer rise faulting following large interplate thrust events. Future seismic failure of the remainder of the seismic gap appears viable, with the northeastern region that has also experienced outer rise compressional activity warranting particular attention. 1. Introduction In September and early October of 2006 several moderate-size (~m b 6.6) and many smaller thrust fault earthquakes occurred near the trench east of the Kuril Islands (Figure 1). About 45 days later, on 15 November, the shallow part of the megathrust failed in an M W 8.3 [global Centroid Moment Tensor (CMT) solution: http://www.globalcmt.org/cmtsearch.html] lowangle thrust-faulting earthquake that ruptured about 200 km northeastward along the trench strike (USGS: 11:14:13.570 UTC, 46.592 N 153.266 E, m b 6.5, M S 7.8). Activity in the outer rise began within minutes of the large megathrust event and continued for the another two months until 13 January 2007, at which time an M W 8.1 (CMT) normal-faulting earthquake ruptured - 2 -

~150 km along a steeply-dipping fault in the outer rise parallel to the trench (USGS: 04:23:21.160UTC, 46.243 N 154.524 E, m b 7.3, M S 8.2). This great earthquake sequence (Figure 2) partially filled a major seismic gap [Kelleher and McCann, 1976; McCann et al., 1979; Lay et al., 1982] northeast of the great 1963 Kuril Islands earthquake [Beck and Ruff, 1987; Kanamori, 1970] and southwest of the great 1952 Kamchatka earthquake [Kanamori, 1976; Johnson and Satake, 1999]. Only modest-amplitude tsunamis were produced around the Pacific by the two earthquakes despite their large size. Tsunami heights less than 1 m were reported for the 15 November 2006 event, except for a 1.76 m wave that caused significant damage in Crescent City, California. The 13 January 2007 event produced tsunami waves less than half the amplitude of the thrust event in most locations, and only 0.37 m high in Crescent City. However, the ground shaking in Japan produced by the January event was significantly stronger than for the November event, as shown in Figure 3. Given the timing of the two events, it is clear that January event was an aftershock, but of unusually large size, qualifying these events as a doublet [Lay and Kanamori, 1980; Kagan and Jackson, 1999]. Activation of moderate size outer rise seismicity after large interplate thrust events has commonly been observed [e.g., Christensen and Ruff, 1988], but the seismic hazard associated with such outer rise aftershock activity has not been considered in detail. Given the characteristically high stress drop and strong short-period seismic wave excitation for intraplate events, this large Kuril doublet provides a rare demonstration of the hazard presented by triggering of very large outer rise events. This intriguing great earthquake sequence also presents an opportunity to examine subduction zone earthquake processes including the relationship of strain accumulation in the upper plate and coseismic slip, large-earthquake interactions within the subducting plate, and - 3 -

stress release and the nature of faulting in the outer rise. Song and Simons [2003] examined gravity variations along many of the world s subduction zones and suggested that large underthrusting events correlate with negative trench-parallel gravity anomalies (TPGA). However, the 15 November 2006 Kuril underthrusting event ruptured a region with a relatively strong positive TPGA, conflicting with the typical pattern. The 2006 rupture is located beneath the only region of the Kuril Islands arc with a significant fore-arc basin. Large seismic slip has been noted to occur elsewhere below fore-arc basins, suggesting an important role for the upper plate in the deformation process [Wells et al., 2003], although large events occur elsewhere along the Kuril zone where there is no fore-arc basin. Outer-rise faulting is common in all subduction zones and is associated with bending stresses acting on the slab just before subduction [e.g. Christensen and Ruff, 1988; Liu and McNally, 1993]. The zone of active shallow normal faulting associated with bending is about 50-100 km wide, and generally represents on the order of 1 Ma of deformation as the lithosphere moves through the region. Normal fault offsets in the outer rise are typically less than ~100 m, so the faults are relatively fresh compared to the megathrust. Christensen and Ruff [1988] suggested that outer-rise faulting is sensitive to stress perturbations associated with large underthrusting events in the adjacent subduction zone. When the interplate zone is frictionally locked, the outer rise may experience shallow compressional events, whereas shallow extensional faulting dominates after the interplate fault ruptures. This interaction may reveal important attributes of the stress state in the subduction zone [Dmowska et al., 1988; Liu and McNally, 1993; Taylor et al., 1996]. Great normal-faulting events in the outer rise like the 13 January 2007 event are relatively rare, and tend to be located seaward of weakly coupled megathrusts [e.g., Kanamori, - 4 -

1971; Lynnes and Lay, 1988], suggesting the possibility of total relaxation of interplate stress during the 15 November 2006 thrust event. The 2006-2007 Kuril Island great doublet events were extensively recorded on broadband global seismic stations. In the following, we analyze this complicated great-earthquake sequence and estimate the slip distribution and other source properties of the two mainshocks using globally recorded surface- and body-wave signals. We extend results of an initial investigation of the sequence [Ammon et al., 2008], using additional methodologies and enhanced procedures to characterize the rupture processes. We then consider the nature of stress interactions between the events. 2. Tectonic Setting Seismic activity and regional tectonic structures in the immediate vicinity of the 2006-2007 Kuril Islands sequence are illustrated in Figure 1. In this region, the Pacific plate subducts at a rate of ~80 mm/yr (NUVEL-1) beneath the North American plate (or a separate Sea of Okhotsk microplate) toward a direction of ~300 N. The Pacific plate lithosphere age at the subduction zone is somewhat uncertain, because it formed during the Cretaceous quiet zone, but is about 100 Ma. There is moderate bathymetric relief on the Pacific plate in the NE Hokkaido rise along the central Kuril island arc, but no profound change in the subducting plate structure [Kelleher and McCann, 1976]. The 2006-2007 earthquake sequence is bordered on the south by a substantial sea-floor canyon (near 46 N, 152.5 E), and is located near the tapered northern end of the obliquely-spread Kuril back-arc basin, which has oceanic crust, whereas surrounding regions of the upper plate have much thicker (~20 km) crust. The broader regional distribution of shallow seismicity along the Kuril island arc prior to the 2006-2007 sequence is shown in Figure 4. Major double-couple mechanisms for CMT - 5 -

solutions from 1977 to 2006 are shown, along with a few additional solutions for earlier large outer rise events from Christensen and Ruff [1988]. Aftershock zones for the 1963 Kuril and 1952 Kamchatka events are outlined, and the occurrence of outer rise extensional faulting seaward of those interplate ruptures is evident. Note that background interplate seismicity defined by the CMT shallow thrust-fault solutions shifts seaward in the vicinity of the 2006-2007 sequence, and appears to be confined to a narrower zone of activity than along the 1963 and 1952 rupture zones. This seaward shift of seismicity in the seismic gap appears to correspond to the region of the fore-arc basin apparent in the bathymetry. There is a corresponding narrowing of the overall width of the trench and a reduction of the free air gravity anomaly along this stretch of the subduction zone [Ammon et al., 2008]. This region also experienced several relatively rare outer rise compressional events, including a particularly large M S 7.2 event on 16 March 1963 [Christensen and Ruff, 1988]. Areas of outer rise compressional activity and extensional activity are highlighted in Figure 4, with one compressional area located in the region of the 2006-2007 sequence and another located near the southwestern end of the 1952 rupture zone. The seismicity time-line in Figure 2 indicates the initial burst of moderate-size thrust events (gray solutions in Figure 1) that commenced in late-september 2006 near the eventual hypocentral region of the 15 November 2006 great event. Waveform modeling of teleseismic P waves for thrust events in the hypocentral region indicates source depths of 10 to 12 km below the ocean bottom, implying a shallow interplate fault dip from the trench of <15. The great thrust event initiated vigorous activity that not only includes aftershocks along a 250 km long region of the interplate boundary region, but comparable simultaneous activity in the outer rise region (light yellow dots in Figure 1). The 13 January 2007 outer-rise event occurred after a two - 6 -

month delay, and was followed by aftershocks (dark yellow dots in Figure 1) that overlap with the earlier outer-rise activity, but are somewhat concentrated to the southeast of the earlier thrustinduced foreshock activity. The megathrust environment had continuing activity after the January event, indicating strong interactions between the two fault zones. The 15 November earthquake decreased substantially after two weeks, gradually decaying to a relatively low level over the next month (Figure 2). The 13 January aftershocks were less numerous in the first two weeks after the mainshock compared with after the 15 November sequence. 3. First-order Attributes of the Great Earthquakes We apply several procedures to characterize the overall geometry and extent of faulting before developing finite-source rupture models for the two great Kuril events. Given the restriction to teleseismic observations of the faulting, and the intrinsically limited resolution or source processes provided by teleseismic signals, we seek constraints on the faulting orientation and rupture velocity that can bound the parameter space of the finite source inversions. W-phase inversion for point-source geometry {Hiroo and Luis need to edit this section;.eps or.pdf versions of the figures are needed so they can be put in final form.} The global CMT solutions for the two mainshocks provide point-source characterizations for body wave and surface wave ground motions ranging from 50 to 200 s in period. For very large earthquakes like these, it is possible that source finiteness effects can bias point-source representations based on this intermediate period range, so we performed inversions of very long-period motions with periods between 200 and 1000 s from the P-wave arrival until the fundamental mode airy phase. The very long-period motions in this time window have been called the W-phase [Kanamori, 1993], and are basically comprised of very - 7 -

long-period, high group velocity spheroidal mode energy superimposed on body wave arrivals. Large ruptures generate sufficient signal-to-noise ratio that energy in this window can be stably inverted for a point-source moment tensor, with accurate prediction of the waveforms being provided by standard Earth models such as PREM [Dziewonski and Anderson, 1981]. Because the waveform interval precedes the main Rayleigh wave arrival, near real-time inversion of the W-phase can provide the most rapid characterization of the very long-period source process for large events for use in tsunami warning systems [e.g., Kanamori and Rivera, 2008], but it is also an intrinsically robust and straightforward way to resolve overall faulting parameters for any large event. W-phase inversion was performed for both mainshocks. For the 15 November 2006 event, 26 broadband stations with particularly stable ultra-long period seismometers were used to invert for a point-source moment tensor. The solution and fits to the data are shown in Figure 5. The best nodal planes for the moment tensor, which a very small intermediate eigenvalue (eigenvalues are 2.04981, -0.05671 and -1.99310 in units of 10 21 Nm), have an orientation of strike φ = 232.1, dip δ = 29.4, rake λ = 115.9, and seismic moment M o = 1.X x 10 21 Nm (M W = 8.14). This is quite similar to the global CMT solution for which the major double couple has orientation φ = 215, δ = 15, λ = 92 and M o = 3.5 x 10 21 Nm (M W = 8.3). Comparison of the fits for the two mechanisms indicate slight improvements for the W-phase inversion, and the strike for this solution is closer to the regional trend of the trench. Moment tensor inversion of intermediate period surface waves alone by J. Polet [personal communication, 2007] gave a similar solution with φ = 206.1, δ = 9.2 and λ =84.3, with M o = 3.9 x 10 21 Nm (M W = 8.3). As apparent in Figure 5, the solution from inversion of the W-phase provides excellent predictions of the Rayleigh wave arrival in the very long-period passband. The dip is not well-resolved, and - 8 -

there is direct trade-off with the seismic moment such that shallower dip results in larger estimated moment (and hence, larger M W ). Given the geometry of the subduction zone and the shallow depth of small events near the hypocenter mentioned above, we prefer a shallower dip, but view the slight differences between the W-phase and CMT solutions as indicative of the uncertainties in the point-source representation, and we will adopt the CMT orientation for the finite rupture models discussed below. Inversion of 20 W-phase signals for the 13 January 2007 event (Figure 6) yielded a solution with moderate intermediate eigenvalue (eigenvalues are 2.39653, -0.35655, -2.03998 in units of 10 21 Nm), and a best nodal plane with orientation φ = 82.6, δ = 20.8, λ = -60.3 and M o = 1.X x 10 21 Nm (M W = 8.16). The CMT solution has orientation φ = 43, δ = 59 and λ = -115, with M o = 1.78 x 10 21 Nm (M W = 8.1). Moment tensor inversion of the intermediate period fundamental mode surface waves alone [J. Polet, 2007, personal communication] had φ = 48.2, δ = 50.7, λ = -103.1, with M o = 1.4 x 10 21 Nm (M W = 8.1). Comparison of synthetics for the various models suggest modest, but systematic differences. All three solutions have one of the nodal planes striking quite obliquely relative to the trench, while the other plane parallels the trench and the trend of the seismicity. The discrepancies between the W-phase and CMT solution is actually larger for this event than for any of the other great earthquakes for which comparisons have been made, suggesting that there is some complexity in the faulting process that is manifested in different moment tensors for different pass-bands. We cannot rule out the possibility that the location of this event near the Kuril trench, where there is steeply dipping bathymetry, causes biases in the solutions based on the 1D PREM Earth model Green functions, but this uncertainty in the overall geometry must be kept in mind for the finite source inversions. - 9 -

Aftershock relocations using double differences While the overall shallow-dipping geometry of the 15 November 2006 event is relatively well-constrained, precise relative location of aftershocks may help to resolve the dip and lateral dimensions of the fault zone. For the 13 January 2007 event, precise aftershock locations might help to distinguish the actual fault plane. Examination of the routine event locations from the National Earthquake Information Center (NEIC) indicated too much scatter to achieve these goals, so we conducted an aftershock re-location procedure using a suite of master events for which we determined accurate source depths by P wave modeling along with double-difference solution for relative locations of clusters of events. {CHUCK and AARON need to provide description and summary figure of results Figure 7 is just a placeholder}. Short-period P-wave back-projection for rupture velocity We apply teleseismic short-period P-wave back-projection using a large aperture seismic array to image rupture front expansion, for the two great events, with the primary goal being to bound the rupture velocity. Short period seismic energy is difficult to model quantitatively due to limitations of Earth models and spatially aliased sampling of the wavefield. However, shortperiod energy can resolve some aspects of overall rupture directivity of large earthquakes using azimuthally distributed stations [e.g., Ni et al., 2005] or spatially dense seismic arrays [e.g., Ishii et al., 2005, 2007; Krüger et al., 2005]. This information can help to constrain lower frequency waveform inversions for which strong trade-offs exist between rupture velocity and spatial slip distribution. We applied back-projection of teleseismic short-period signals recorded across a large network of broadband sensors in western North America at source-receiver distances of ~50-80 - 10 -

from the Kuril doublet events. A two-dimensional spatial grid at the surface encompassing the source region was defined, and differential travel time predictions for an assumed reference velocity model (IASPEI91) were used to shift and sum all observations as though they originated at each grid point at corresponding rupture times. The change in travel time curve derivative (dp/dδ) and spatial distribution of the receivers yield space-time isolation of loci of coherent high frequency radiation from the fault from which average rupture velocity and source directivity can be measured as well as lower bounds for rupture length and duration. This method has very little depth resolution since relative differences of dp/dδ do not vary significantly across tens of kilometers of depth. This short-period back-projection method does not recover absolute slip, and is primarily sensitive to large and localized bursts of energy release rather than to energy released simultaneously over spatially extended rupture fronts, so the images only characterize some aspects of the source radiation. However, rupture velocity can be inferred from the space-time pattern of any high frequency energy release bursts, and this is particularly difficult to resolve otherwise. Alignment of the short-period signals is important, and difficult, and time averaging over the stacked signals is needed to deal with both imprecise alignment of signals and the fact that the energy used is beyond the source corner frequency and intrinsically involves interference effects due to spatial and temporal finiteness. Station statics for the emergent 15 November 2006 event were obtained by picking relative P wave arrival times for a large thrust pre-shock on 10 October 2006, close to the November epicenter, while the first arrivals were directly aligned for the more impulsive 13 January 2007 event. A 30 s time integration was applied in the final back-projections in order to spatially isolate bursts of coherent short-period radiation received across the continental array. - 11 -

We use band-pass filtered (0.8 2.0 Hz) P-wave data from approximately 280 stations in North America for the 15 November 2006 event and approximately 350 stations for the 13 January 2007 event. Many of the stations are from the Transportable Array of USArray. We also performed back-projections for a smaller number (~70) of broadband stations located in Europe and for the Hi-Net array in Japan, but the distance ranges involved were not as suitable for imaging the source processes for the Kuril doublet as was the case for the North American data. The number and density of stations of Hi-Net is ideal for back-projection applications for different regions such as the 2004 Sumatra [Ishii et al., 2004] and 2005 Nias [Walker et al., 2005] earthquakes, but the distances to the Kuril events span the upper mantle triplication range from ~18-30 degrees. This makes predicting the correct dp/dδ difficult, and there is rangevarying waveform distortion associated with the triplicated arrivals [Walck, 1984], making backprojection of the wavefield unstable. Back-projection images for the European data contained more smearing, both spatially and temporally, relative to the images using North American data because the array footprint is much smaller. The European data are also at further distances (~75-90 degrees) where the incoming wave field has a smaller dp/dδ, reducing space-time isolation of energy bursts, so we only present the results for North American stations. The back-projections use nth-root stacking, with some variation in image fidelity depending on the power (n) used. Areas from which the energy is radiated show up as bright spots in the grid for the time step at which that part of the fault was radiating P-waves at periods of about 1 Hz. Finite array aperture produces time-varying streaking of the stacked images in the direction toward the array. Our grid is evenly spaced at 0.15. The North American array results are shown in Figure 8. The top panels show energy radiation at three time steps, one near the rupture initiation and two images from later parts of each rupture. For both events, the ruptures - 12 -

are dominated by a northeastward component of the rupture, with distinct bursts of energy offset in space and time. The corresponding full time sequence of the back-projections for each event are available as movies provided in on-line supplements. The approximate short-period time functions given by the peak amplitude of the image over the entire grid formed at each time step are shown for both events at the bottom of Figure 8. The limited aperture smears the time resolution, and these radiation histories are only appropriate for the azimuth toward North America. In order to suppress the imaging artifacts intrinsic to this method, we deconvolve the array response by the corresponding space-time image formed for a smaller reference earthquake. For a reference event we chose the M w 6.5 1 October 2006 pre-shock, which was large enough to generate high signal-to-noise ratio first arrivals. With a rupture length probably between 10-20 km, this event is small enough to approximate a point source for which the backprojection smearing in space-time is controlled by the array response, or point-spread function. This event was used to remove the array response for both doublet events (Figure 9). In each case, only stations that recorded high signal-to-noise ratio P-wave arrivals for both the reference and main events were used. We approximate the three-dimensional (latitude, longitude and time) deconvolution by iteratively subtracting the reference event images for a large time window surrounding the event. We align the events using the space and time shifts that yield the maximum three-dimensional cross correlation coefficient between the reference and main event. For each iteration, the values of the reference event being subtracted are weighted by the reciprocal of the cross correlation coefficient, which effectively terminates the procedure when the maximum cross correlation coefficient is low. The result allows us to identify the time and location corresponding to regions - 13 -

of significant bursts of high frequency energy, which are treated as individual subevents. This procedure objectively defines the locations and times of the subevents apparent in the backprojection images like Figure 8. We obtain lower bounds of rupture length and duration using the subevents with the highest cross-correlation coefficients, of approximately 155 km and 83 s for the 15 November 2006 event and 130 km and 33 s for the 13 January 2007 event. If we infer smooth rupture between the peak bursts for the two subevents with the highest cross correlation coefficients in the images, we obtain a rupture velocity of about 2.0 km/s for the 15 November 2006 event and 3.8 km/s for the 13 January 2007 event. The values vary slightly depending on the power (n) used (Figure 9). This short-period rupture imaging is predominantly sensitive to local concentrations of high frequency radiation, and may not sense the smoother slip processes in the rupture. To explore the smoother rupture process, we must also consider the wavefield sensitive to the lowerfrequency processes. Surface wave source time function directivity The Kuril Islands doublet events each provided several hundred global broadband Rayleigh wave recordings. Since surface-wave phase velocities are close to typical earthquake rupture speeds, the observed Rayleigh waveforms often contain large directivity effects [e.g. Ammon et al., 2003; Ammon et al., 2006]. Deconvolution of propagation operators can isolate the azimuthally dependent effective source time functions (STFs). The STFs can be used to resolve the rupture length, characterize the propagation, and image the smooth components of the seismic moment distribution. In previous analyses of large events we used small earthquakes with similar location and fault orientation to the mainshock to provide empirical Green s functions (EGFs) [e.g., Ammon et al., 1993; Velasco et al., 2000]. One limitation of EGFs are the - 14 -

difficulties of reliably isolating very long period (> 250 s) components of the mainshock due to the intrinsically weak long-period excitation of the smaller events. To circumvent this limitation Ammon et al. [2006a, 2006b] used point-source synthetic seismograms (theoretical Green s functions: TGFs) computed using normal-mode summation (down to periods of 20 s) for the PREM [Dziewonski and Anderson, 1981]. For large or slow ruptures the TGF analysis allows one to resolve directivity. For shorter overall rupture durations, the low-frequency restrictions of the TGF analysis blurs some important directivity patterns that are only prominent at the shorter periods. Material needed from CHUCK: Discuss correction of TGF by aspherical structure. Results of simple directivity analysis for 15 November 2006 and 13 January 2007, giving basic length, rupture velocity constraints. 4. Finite Fault Rupture Images of The Great Events Guided by the constraints from the preceding analyses, we developed finite-fault slip models for the two great events. We initially inverted large P and SH data sets for each event using flexible finite-source algorithms that allow for changes in subevent moment tensor with unconstrained or constrained rupture velocities for variable hypocentral depths and fault dimensions [Kikuchi and Kanamori, 19xx]. This suite of inversions convinced us that the body waves provide no indication (or resolution) of change in fault orientation during rupture for either event, and that the CMT and other point-source solutions discussed above all provide adequate basic geometries for matching the waveforms for both events. Two algorithms are applied to produce final slip models, both prescribing the fault geometry and subdividing it into a - 15 -

grid of subfaults, but then imposing different constraints on the rupture and using slightly different data sets. The first algorithm specifies a constant rupture velocity from the hypocenter and allows distinct source time functions for each subfault element, with a linear inversion of teleseismic P and SH observations. The rake is allowed to vary at each subfault. Green functions were computed for simple layered source and receiver structures connected by geometric spreading for a deeper JB Earth model. The second method is a search-based algorithm [Velasco et al., 2000; Ammon et al., 2006a,b] that estimates a smooth seismic moment distribution that matches observed teleseismic P and SH waveforms and R1 STFs in a least-squares sense. In the local search algorithm, the rupture model is successively perturbed in a search for better fitting models. A specified constant rupture velocity is used. We use a zero-slip initial model and include several thousand perturbations. The time function of each subfault is a half cosine function like that used by Ji et al. [2002], with a duration parameter that was allowed to vary among 4, 8, 12, 16 or 20 s. The overall mechanism is held fixed, with constant rake over the fault plane. For the R1 STFs, the mechanism and depth of each point source in the model are assumed to be identical to those of the point-source specified in the STF estimation. Numerical experiments suggest that the main limitations from ignoring depth variations in the STF computations are relatively small amplitude, low frequency artifacts in the STF estimates which are not of major concern for estimating first-order rupture characteristics [e.g., Ammon et al., 2006a,b]. For body wave computations, the depth is adjusted along the dipping fault plane. The body wave data are corrected for relative geometric spreading to a uniform distance of 60. Both inversions assume frequency-independent attenuation models with t* = 1 s for P-waves and t* = - 16 -

4 s for S-waves, and determine subevent point-source seismic moment, which we convert to slip using the grid spacing dimensions and the local shear modulus. Body-Wave Selection and Processing We used body-waves from distant stations to minimize PP and SS interference. We performed inversions of data with and without deconvolution of the individual instrument responses. Instrument deconvolution tends to enhance long-period noise, causing baseline uncertainties. Concentrating on signals recorded by STS-1 instruments, we found that simple integration of the raw seismograms gave stable estimates of ground displacement over the passband emphasized in our inversions, so we preferred that procedure for the body wave only inversions. For the search-based inversions we band-pass filtered the instrument-deconvolved body waves to include periods between 150 and 3 s, relying on the R1 STFs to resolve the lower frequency component of the source spectrum. We do not account for core reflections, which are generally of small amplitudes; where possible we checked this by examining the waveforms from smaller events in the sequence. We selected subsets of the huge available global data set to balanced azimuthal weighting in the inversions. Relative timing is of great importance for finite source inversions using body waves, because small differential times within the waveforms provide spatial resolution. Both great earthquakes have somewhat emergent initial P-wave and SH-wave onsets, thus travel time alignment is not trivial. In order to reduce the subjectivity of onset picks, we used the NEIC hypocentral locations and computed travel times for the PREM Earth model, applying aspherical path corrections for a P-wave tomographic model [Christine Houser, personal communication, 2007] and an SH-wave tomographic model [Megnin and Romanowicz, 2000]. This gave very - 17 -

good relative alignments consistent with direct picks of the more impulsive arrivals. SH waves were used for azimuths away from nodes. Slight adjustments in SH onset times for a few stations were made after initial inversions established that the entire waveform was slightly shifted. The 15 November, 2006 Fault Rupture For the finite source inversions for the 15 November 2006 event, we adopted the CMT major double couple fault geometry with φ = 214, δ = 15, and for fixed rake inversion, λ = 92. This shallow dipping plane is consistent with the distribution of aftershocks inboard of the trench, although the strike is slightly oblique to the trench axis (it is consistent with broader regional seismicity). Very similar slip distributions are found when the strike is chosen to parallel the trench, but variance reductions do not guide selection of a mechanism different than the CMT solution. A large, well-distributed set of P-waves was inverted with the variable rake algorithm. Figure 11 shows the slip model, source time function, and waveform matches obtained for a rupture velocity of 2.0 km/s, as indicated by the short-period back-projection results. The source velocity structure included a 3 km deep ocean layer overlying a 15 km thick crust with V p = 6.5 km/s, V s = 3.74 km/s and ρ = 2.87 gm/cm 3, and a mantle layer with V p = 7.8 km/s, V s = 4.4 km/s and ρ = 3.3 gm/cm 3. The hypocentral depth was set to 15 km (12 km deep into the crust), based on the modeling of nearby aftershocks. The subfault grid had 20 km spacing along strike and 10 km spacing along the dip direction. Each subfault source time function was parameterized by 7 overlapping 6 s duration triangles offset by 3 s each. The slip is found to be concentrated in the upper portion of the rupture plane, with three regions of highest slip along the strike. The slip extends along about 220 km, with only modest variations in the rake over the fault plane. The - 18 -

average rake is 96, very similar to the CMT solution. The source time function has a total duration of about 110 s, with three concentrations of moment release and a total seismic moment of 3.8 x 10 21 NM (M W = 8.3). The rupture extends toward the NE, consistent with the overall R1 STF directivity in Figure 10. Comparison with the positions of the high frequency bursts of energy imaged by the short-period back-projection (Figure 7) suggests that these bursts are associated with the two northeastern patch of enhanced slip. The patch of slip near the hypocenter may have failed bilaterally, thereby not giving a strong high frequency feature at the North American azimuth, whereas the rupture would have been progressing northeastward as it encountered the other two patches, accounting for the enhanced bursts of short-period radiation toward North America. The shallow slip for this model is as large as 11.4 m, which is somewhat difficult to reconcile with the lack of tsunami excitation. Inversions for varying rupture velocities and varying fault dimensions recover very similar basic features, but with strong tradeoffs in rupture velocity and spatial location of slip and comparable variance reduction. Assumption of a deeper hypocenter somewhat shifts slip to greater depth for the first subevent, but the models clearly favor shallow slip to match the P waves. THIS PART AND FIGURE 12 NEEDS TO BE UPDATED BY CHUCK FOR SOLUTION WITH STF CORRECTED FOR ASPHERICAL STRUCTURE. Joint inversion of P, SH and R1 STF observations was performed using the search algorithm, holding the rake fixed at the CMT value (92 ). For these inversion, we located the top of the rupture plane at 1 km beneath the trench. The source velocity model is a water layer over a half-space with V p = 6.7 km/s and V s = 3.87 km/s. Figure 12a shows a slip model obtained assuming a rupture velocity of 2.0 km/s. Slip is computed from the subfault moment assuming a shear-modulus of 40 GPa. The moment-rate history is similar to that for the perhaps the best-resolved feature of - 19 -

the rupture and is relatively invariant with assumptions about rupture speed, and robust to changes in the waveforms used in the inversion. As for the variable rake inversion, there are 3 intervals of higher moment release over the 120 s source duration. The inclusion of the longperiod R1 information does not significantly change the moment rate function, but the slip appears to be spatially more slowly varying than for the P-wave inversion. Good matches are obtained to the set of body and surface waves used in this inversion (Figure 12b). The largest slip was located about 25-100 km from the hypocenter (at a depth of 11 km), on the margin of a large slip region. For a fixed rupture speed, the location peak slip areas are better constrained along strike than along dip; in sensitivity tests of the inversion we found that they can be moved around a few 10 s of km, and some high-wavenumber characteristics of the large, deep asperity are difficult to constrain. The seismic moment for this model is 4.6 x 10 21 Nm (M w = 8.4). The average slip for this model is about 4.6 m. The 13 January, 2007 Fault Rupture The CMT faulting geometry for this event (φ = 43, δ = 59, λ = -115 ) is rather unusual (Figure 1), suggesting oblique slip on either a nearly east-west striking plane, or along a plane roughly parallel to the trench and the foreshocks and aftershocks (Figure 1). As shown above the seismicity relocations do not uniquely resolve the issue, and the W-phase inversion suggests the possibility of some complexity in the fault orientation. Finite source models for the 13 January 2007 event were obtained for the CMT, W-phase and Polet (and many other) fault geometries dipping either southeast or northwest in an effort to resolve the fault plane ambiguity. Similar slip distributions and overall waveform fits are found for all of the fault orientations, with concentration of slip in the upper 25 km of the oceanic lithosphere, and little independent - 20 -

constraint on rupture velocity or precise geometry. For lower rupture velocities near 2 km/s the estimated peak slip at crustal depths is larger, over fault lengths of 70 to 90-km. Larger rupture velocities of up to 4.0 km/s yield estimated peak slip of 10-20 m and total rupture lengths of 120-160 km. The weak tsunami excitation tends to favor lower slip, and corresponding higher rupture velocities, the latter being compatible with the estimates of rupture velocity of 3.8 km/s from the short-period back projection (Figure 7). The fit to P waveforms at stations to the south tends to be better for finite source models using the northwest dipping plane, but stations to the east are fit slightly better for the finite source models using the southeast dipping geometry. While prescribing a strike for either plane parallel to the aftershock distribution results in a good fit to the P waves, the long-period Love and Rayleigh wave signals together require some oblique slip overall, and we ultimately adopt the CMT orientation to ensure compatibility with long-periods. While slightly better overall fits were obtained for the northwest dipping fault geometry if we assume a shallow hypocenter, this fault plane deviates from the trend of the aftershocks and trench. It is possible that aftershocks are an unreliable guide as to the fault geometry, given that their mechanisms deviate from the great event (Figure 1), but we chose the southeast dipping plane of the CMT solution for the final model since this does have a trend compatible with the aftershocks. Figure 13a shows the slip distribution obtained from fitting P and SH waves with the variable rake algorithm using the southeast dipping fault in the CMT solution, for a specified rupture velocity of 3.8 km/s, based on the back-projection results in Figure 7. The source velocity model in this case had a 4 km deep ocean over a 6 km thick crust with V p = 6.8 km/s, V s = 3.92 km/s and ρ = 2.7 gm/cm 3, and a mantle with V p = 7.9 km/s, V s = 4.56 km/s and ρ = 3.3 gm/cm 3. The rake is found to be fairly stable, with an average value of -111 relative to the - 21 -

CMT solution of -115. For this case, having the hypocenter at a depth of about 23 km gives a better solution than a shallow hypocenter as it allows some northwestward propagation of the rupture. The moment rate function has one main pulse with a broadened shoulder that is enhanced by inclusion of the SH waves, giving rise to two regions of peak slip spaced about 110 km apart. These two areas of peak slip appear to correspond to the two primary high frequency bursts in the short-period back-projection image in Figure 7. The subfault grid spacing was 21.7 km along strike and 10.9 km along the dip direction. The subfault source time functions were parameterized by 6 overlapping 6 s duration triangles shifted by 3 s each. The slip is largest within the crustal layer, with slip as much as 12.5 m for this solution. The overall seismic moment estimate is 2.1 x 10 21 Nm (M W = 8.1). The rupture has a duration of at least 60 s, and the early part of the waveforms are quite well fit (Figure 12b), but there does appear to be coherent energy later in the body waves. For a high rupture velocity simply extending the fault dimensions leads to late, spatially poorly-resolved slip far along strike to the northeast, beyond the aftershock zone. This suggests that a uniform rupture velocity may not be correct for this rupture, and indeed inversions that do not prescribe a rupture velocity find better fits to the data with some later, somewhat deeper slip on the fault. THIS PART AND FIGURE 14 NEEDS TO BE UPDATED BY CHUCK FOR SOLUTION WITH STF CORRECTED FOR ASPHERICAL STRUCTURE AND RUPTURE VELOCITY OF 3.8 KM/S. A slip model obtained simultaneous inversion of P, SH and R1 STFs, with a fixed rake given by the CMT solution and a rupture velocity of 3.5 km/s is shown in Figure 14a. The source region half-space velocity structure has V p = 7.25 km/s and V s = 4.18 km/s, respectively. The waveform fits are shown in Figure 14b. The rupture is again characterized by an ansymmetry toward the northeast with primary slip located in the upper 25-22 -

km of the fault, over a length of about 175 km. The seismic moment is 1.5 x 10 21 N-m (M W = 8.1). The average slip is about 9.6 m over a 120 km 20 km fault extent (converting from moment to slip using a shear modulus of 52 GPa). Some weaker slip appears to extend to depths of 30-35 km, but resolution is limited. The moment rate function is again dominated by a large pulse with a secondary shoulder and a duration of 40 s, with some weak energy release during the next 20 s. This latter energy is not well accounted for by kinematic rupture models with high constant rupture velocity, but appears to come from relatively large depth (>25 km) not too far from the epicenter. Delayed, deep slip on the fault is a tentative explanation for this energy, but more complex rupture parameterizations will need to be examined to resolve it better. The 13 January 2007 (M W = 8.1) event is the third largest outer rise normal faulting event recorded, after the 1933 Sanriku, Japan (M W = 8.4) [Kanamori, 1971] and 1977 Sumba, Indonesia (M W = 8.3) [Lynnes and Lay, 1988] earthquakes. Source spectra The dramatically different moment rate functions and source spectra for the two great events are shown in Fig. 15. The body wave spectral amplitudes for the 13 January 2007 event are significantly larger, by ratios of 4 to 7, than those for the 15 November 2006 event, despite the larger seismic moment of the earlier event. The January event thus has larger 1 s period body wave magnitude (m b ) and 20 s period surface wave magnitude (M s ) than the November event. This short-period enrichment is similar to that for the 1933 Sanriku earthquake [Kanamori, 1971], and may reflect rupture on a fault with little cumulative slip. Seismic energy release for the November event (9.6 x 10 15 J) is less than for the January event (4.3 x 10 16 J), and the energymoment ratios are 2.7 x 10-6 and 2.4 x 10-5, respectively 12. The factor of 15 contrast in scaled - 23 -

energy indicates significant differences between interplate and intraplate faulting environments. Triggering of a large outer rise rupture with strong high frequency shaking constitutes an important potential seismic hazard that needs to be considered in other regions. 5. Discussion KEVIN TO GIVE RESULTS AND A FIGURE (?) SHOWING CALCULATION OF STRESS CHANGES The slip models from our body and surface wave joint inversions are shown in map view in Figure 16. The last few decades of outer rise activity along the Kuril Islands arc primarily involved extensional faulting seaward of earlier large interplate thrusts, with the exception of several outer rise compressional events distributed along the former seismic gap (Figure 4). These include a large M S = 7.2 event on 16 March 1963 (46.79 N, 154.83 N), located in the outer rise offshore of the 15 November 2006 rupture zone (Fig. 16). The transition from outer rise compression to outer rise tension following the 2006 interplate thrust supports the notion of outer rise stress modulation by varying interplate frictional stresses [Christensen and Ruff, 1988; Lay et al., 1989; Lin and Stein, 2004; Taylor et al. 1996]. The Kurile doublet provides the clearest example of the full temporal pattern through the seismic cycle yet observed. The January outer rise event was unusually large, comparable to the great extensional events in uncoupled seismic zones, suggesting that the November event completely relaxed friction on the megathrust, allowing slab-pull forces to operate unimpeded on the outer rise. Stress transfer occurred on multiple time scales. Initial outer rise activity commenced within 40 minutes of the large thrust event, suggesting that dynamic and/or static stress transfer triggered events in a highly strained segment of the Pacific Plate. The 60 day delay before the second member of the - 24 -

doublet indicates a longer response time, consistent with a visco-elastic strain migration rate of 100-km/60-days, comparable to that for the 30 March 1965 M w = 7.7 outer rise earthquake triggered by the great 4 February 1965 Rat Island earthquake [Melosh, 1976]. 6. Conclusions The 2006-2007 Great Kuril earthquake sequence involved both underthrusting and extensional faulting on a large scale which has not been previously observed. The great thrust event of 15 November 2006 ruptured the shallow region of the megathrust fault between the Pacific plate and the Kuril arc, with 6-10 m of slip along a 200 km long segment of the boundary. This produced extensive aftershock activity along the megathrust as well as in the outer rise region, and two months later the 13 January 2007 normal-faulting event released comparable amounts of slip on a steeply dipping intraplate fault near the trench. The outer rise had previously experience relatively large compressional events, so it appears that the shallow stress environment cycled from compression to extension with the stress release on the megathrust. Acknowledgements. We thank the developers of GMT and SAC, and the seismic network operators who have constructed a superb open-data-access network for seismic research and monitoring. Ed Garnero and Christine Houser provided algorithms used for computing aspherical model travel time corrections for tomography models. The facilities of the IRIS Data Management System were used to access the data used in this study. Supported by NSF grants EAR0453884 and EAR0635570 (TL) and USGS Award Number 05HQGR0174 (CJA). - 25 -

References Ammon, C. J., A. A. Velasco, and T. Lay (1993), Rapid estimation of rupture directivity: Application to the 1992 Landers (M S = 7.4) and Cape Mendocino (M S = 7.2) California Earthquakes, Geophys. Res. Lett., 70, 97-100. Ammon, C. J., A. A. Velasco, and T. Lay (2006a), Rapid estimation of first-order rupture characteristics for large earthquakes using surface waves: 2004 Sumatra-Andaman earthquake, Geophys. Res. Lett., 33, L14314, doi:10.1029/2006gl026303, 2006a. Ammon, C. J., H. Kanamori, T. Lay, and A. A. Velasco (2006b), The 17 July 2006 Java tsunami earthquake (M w = 7.8), Geophys. Res. Lett., 33, L24308, doi:10.1029/2006gl028005. Ammon, C. J., H. Kanamori, and T. Lay (2008), A great earthquake doublet and seismic stress transfer cycle in the central Kuril islands, Nature, in press. Beck, S. L., and L. J. Ruff (1987), Rupture process of the great 1963 Kurile Islands earthquake sequence: Asperity interaction and multiple event rupture, J. Geophys. Res., 92, 14123-14138. Christensen, D. H., and L. J. Ruff (1988), Seismic coupling and outer rise earthquakes, J. Geophys. Res., 93, 13,421-13,444. Dmowska, R., J. R. Rice, L. C. Lovison, and D. Josell (1988), Stress transfer and seismic phenomena in coupled subduction zones during the earthquake cycle, J. Geophys. Res., 93, 7869-7884. Dziewonski, A.M., and D.L. Anderson (1981), Preliminary reference Earth model, Phys. Earth Planet. Int., 25, 297-356, 1981. Ishii, M., P. M. Shearer, H. Houston, and J. E. Vidale (2005), Extent, duration and speed of the 2004 Sumatra-Andaman earthquake imaged by the Hi-net array, Nature, 435, 933 936. - 26 -

Ishii, M., P. M. Shearer, H. Houston, and J. E. Vidale (2007), Teleseismic P wave imaging of the 26 December 2004 Sumatra-Andaman and 28 March 2005 Sumatra earthquake ruptures using the Hi-net array, J. Geophys. Res. 112, B11307, doi:10.1029/2006jb004700. Ji, C., Wald, D. and Helmberger, D. V. (2002), Source description of the 1999 Hector Mine, California, earthquake; Part I, Wavelet domain inversion theory and resolution analysis, Bulletin of the Seismological Society of America, 92, 1192-1207. Johnson, J. M., and K. Satake (1999), Asperity distribution of the 1952 great Kamchatka earthquake and its relation to future earthquake potential in Kamchatka, Pure Appl. Geophys., 154, 541 553. Kagan, Y. Y, and D. D. Jackson (1999), Worldwide doublets of large shallow earthquakes, Bull. Seismo. Soc. Am., 89, 1147-1155. Kanamori, H. (1970), Synthesis of long-period surface waves and its application to earthquake source studies-kurile Islands earthquake of October 13, 1963, J. Geophys. Res., 75, 5011-5027. Kanamori, H. (1971), Seismological evidence for a lithospheric normal faulting; the Sanriku earthquake of 1933, Phys. Earth Planet. Inter., 4, 289-300. Kanamori, H. (1976), Re-examination of the earth's free oscillations excited by the Kamchatka earthquake of November 4, 1952, Phys. Earth Planet. Inter., 11, 216-226. Kanamori, H. (1993), W-phase, Geophys. Res. Lett., 20, 1691-1694. Kelleher, J., and W. McCann (1976), Buoyant zones, great earthquakes and unstable boundaries of subduction, J. Geophys. Res., 81, 4885-4896. Kirby, S. H., E. R. Engdahl, and R. Denlinger (1996), Intraslab earthquakes and arc volcanism: Dual physical expressions of crustal and uppermost mantle metamorphism in subducting slabs, in Subduction: Top to Bottom, edited by G. E. Bebout, et al., pp. 195 214, American Geophysical Union. - 27 -

Krüger, F., Ohrnberger, M., (2005), Tracking the rupture of the Mw=9.3 Sumatra earthquake over 1,150 km at teleseismic distance, Nature, 435, 937-939. Lay, T., and H. Kanamori (1980), Earthquake doublets in the Solomon Islands, Phys. Earth and Planet. Inter., 21, 283-304. Lynnes, C. S., and T. Lay (1988), Source process of the great 1977 Sumba earthquake, J. Geophys. Res., 93, 13,407-13,420. Lay, T., H. Kanamori, and L. Ruff (1982), The asperity model and the nature of large subduction zone earthquakes, Earthquake Pred. Res., 1, 3 71. Lay, T., L. Astiz, H. Kanamori, and D. H. Christensen (1989), Temporal variation of large intraplate earthquakes in coupled subduction zones, Phys. Earth Planet. Inter., 54, 258-312. Lin, J., and R. S. Stein (2004), Stress triggering in thrust and subduction earthquakes and stress interaction between the southern San Andreas and nearby thrust and strike-slip faults. J. Geophys. Res. 109, B02303, doi:10.1029/2003jb002607. Liu, X., and K. C. McNally (1993), Quantitative estimates of interplate coupling inferred from outer rise earthquakes, Pure Appl. Geophys., 140, 211-255. Lynnes, C. S., and T. Lay (1988), Source process of the great 1977 Sumba earthquake, J. Geophys. Res., 923, 13,407-13,420. McCann, W.R., S.P. Nishenko, L.R. Sykes, and J. Krause (1979), Seismic gaps and plate tectonics: seismic potential for major boundaries, Pure Appl. Geophys., 117, 1082-1147. Mégnin, C., and B. Romanowicz (2000), The three-dimensional shear velocity structure of the mantle from the inversion of body, surface, and higher-mode waveforms, Geophys. J. Int., 143, 709-728, 2000. Melosh, H. J. (1976), Nonlinear stress propagation in the Earth s upper mantle, J. Geophys. Res., 81, 5621-5632. - 28 -

Song, T. A., and M. Simons (2003), Large trench-parallel gravity variations predict seismogenic behavior in subduction zones, Science, 301, 630-633. Taylor, M. A. J., G. Zheng, J. R. Rice, W. D. Stuart, and R. Dmowska (1996), Cyclic stressing and seismicity at strong coupled subduction zones, J. Geophys. Res., 101, 8363-8381. Velasco, A. A., C. J. Ammon, and T. Lay (1994), Empirical Green function deconvolution of broadband surface waves: Rupture directivity of the 1992 Landers, California (M w = 7.3), earthquake, Bull. Seism. Soc. Am., 84, 735-750. Velasco, A. A., C J. Ammon, and S. L. Beck (2000), Broadband source modeling of the November 8, 1997, Tibet (M w = 7.5) earthquake and its tectonic implications, J. Geophys. Res., 105, 28065-28080. Venkataraman, A., and H. Kanamori (2004), Observational constraints on the fracture energy of subduction zone earthquakes, J. Geophys. Res., 109, 10.1029/2003JB002549. Walck, M.C. (1984), The P-wave upper mantle structure beneath an active spreading center; the Gulf of California, Geophys. J. Roy. Astron. Soc., 76, 697-723. Waldhauser, F., and D. Schaff (2007), Regional and teleseismic double-difference earthquake relocation using waveform cross-correlation and global bulletin data, J. Geophys. Res., 112, B12301, doi:10.1029/2007jb004938. Walker, K.T., Ishii, M., and Shearer, P.M. (2005), Rupture details of the 28 March 2005 Sumatra M w 8.6 earthquake imaged with teleseismic P waves, Geophys. Res. Lett., 32, L24303, doi: 10.1029/2005GL024395. Wells, R. E., R. J. Blakely, Y. Sugiyama, D. W. Scholl, and P. A. Dinterman (2003), Basincentered asperities in great subduction zone earthquakes; a link between slip, subsidence and subduction erosion? J. Geophys. Res., 108, 2507, doi:10.1029/2002jb002072. - 29 -

Figure 1. Map showing the global CMT solutions for the larger earthquakes in the 2006-2007 Kuril Islands sequence (focal mechanisms), NEIC epicenters of the largest events (large circles), activity between the two largest events (yellow circles), aftershocks following the 13 January 2007 event (orange circles), and prior activity (gray circles). The gray shaded focal mechanisms identify foreshocks of the 15 November 2006 event, the red-shaded mechanisms occurred after the 15 November 2006 event. Focal mechanisms are plotted at the NEIC epicenters. The stars indicate the CMT centroid locations for the two mainshocks, which are shifted seaward relative to the NEIC locations. - 30 -

Figure 2. Space-time seismicity diagram showing the temporal evolution of the 2006-2007 Kuril Islands earthquake sequence. The September foreshock and the two main shock sequences are clear in the chart, which shows earthquake activity as a function of time and distance along the trench (the 15 November earthquake origin time and hypocenter are used as the reference). The arrow shows the estimated plate motion computed using model NUVEL-1 with North-America fixed. - 31 -

Figure 3. Peak ground acceleration amplitudes recorded at K-NET stations in Japan, indicating the significantly stronger shaking produced by the 13 January 2007 event. - 32 -

Figure 4. NEIC shallow seismicity distribution and all CMT solutions for events along the central Kuril Island region prior to the 15 November 2006 event. CMT centroids have an overall bias to the southeast. The approximate along-strike lengths of the 1963 Kuril Islands and 1952 Kamchatka earthquakes and the 2006-2007 great doublet are shown by red lines. Outer rise activity of extensional or compressional nature is highlighted. The outer rise compressional mechanisms in red are from Christensen and Ruff [1988]. - 33 -

Figure 5. W-phase waveform inversion results and waveform fits for November 15, 2006. - 34 -

Figure 6. W-phase inversion results for January 13, 2007. - 35 -

Figure 7. Double-difference relocation results. - 36 -

Figure 8. Images of the great doublet ruptures formed by short-period P-wave back-projection using signals from ~300 broadband seismic stations in western North America. The top row shows the 0.8-2.0 Hz P-wave energy stacked and integrated over 30 s time windows centered on three time points for the 15 November 2006 rupture, the second row is the same for the 13 January 2007 rupture. Warm colors identify regions of high short-period P-wave energy radiation bursts relative to adjacent regions of low-amplitude or no P-wave radiation (blue colors). The star in each image is the NEIC epicenter of the corresponding event. The time varying peak-amplitude of stacked signal over the back-projection grid is shown below for each event, with the time points corresponding to the images above indicated by the triangles. These peak amplitude traces are crude approximations of the short-period source time function corresponding to the direction to North America, but they are affected by the limited space-time isolation of radiation resulting from the finite array aperture. - 37 -

Figure 7. Short-period energy bursts and inferred rupture velocities for the Kuril doublet events (top row 15 November 2006; bottom row 13 January 2007) determined by iterative deconvolution of the short-period back-projection images by point-spread functions determined for identical receiver configurations using the 01 October 2006 M w = 6.5 pre-shock. The circle locations are at positions of maximum cross correlation coefficient (ccc) for each iteration, with the numbers indicating relative time. The circle diameters are scaled by the (ccc), and indicate relative amplitude of the subevents. The rupture velocities shown at the lower right of each box are based on the two largest subevents, assuming continuous rupture propagation between them. Back-projection stack values are calculated using linear stacking in the left column, while the results in the right column are based on values using cube-root stacking. Linear stacking is more sensitive to individual seismogram amplitudes, while cube-root stacking is almost entirely sensitive to the coherency of an arrival across the receiving array. The reference event used was close to the epicenter of the November event and had a similar focal mechanism to it. - 38 -

Figure 10. Directivity pattern observed in R1 STF estimates computed using TGF with aspherical Earth corrections for 15 November 2006. Assuming a simple unilateral rupture, the linear increase in duration with directivity parameter, gamma, is consistent with a 120s long duration and a fault length of 215 km. These correspond to a rupture speed about equal to 1.8 km/s. - 39 -

Figure 11a. Finite source model for the 15 November 15 2006, from inversion of only teleseismic P waves. The fault plane orientation is for the NW-dipping plane from the GCTM major double-couple solution (top right), and has a strike of 214 and a dip of 15. The hypocentral depth was set to 15 km, slip on a 12 x 10 km grid with 20 km spacing along strike and 10 km spacing down dip was determined assuming a rupture velocity of 2 km/s. The rake for each subfault was allowed to vary, and the slip magnitudes and directions are indicated on the left. The source time function is shown on the right, and the seismic moment estimate of 3.8 x 1021 Nm gives an Mw of 8.32. - 40 -

Figure 11b. P-wave inversion waveform fits for Vr 2.0 km/s for November 15, 2006. - 41 -

Figure 12a. Slip model obtained from joint inversion of P-, S- and R1-waves and moment-rate history of the 15 November 2006 M W 8.3 earthquake. Moment was converted to slip assuming a shear modulus of 40 GPa. The circles are rupture isochrones for the assumed rupture speed of 2.0 km/s. The moment-rate history is aligned with the slip map to illustrate the correspondence between the two views of the rupture. - 42 -

Figure 12b. (Top) distributions of data used in the inversion. R1 radiation pattern at 100 s is shown in the STF azimuthal map. (Bottom) Waveform fits corresponding to the model shown in Figure 13a. The black line identifies observations, the red line, predictions. Signal type and distance in degrees are shown above and to the left, station and azimuth are shown above and to the right of each waveform. The model fits about 85% of the signal power. - 43 -

Figure 13a. P and SH wave inversion results for January 13, 2007. - 44 -

Figure 13b. P and SH wave inversion waveform fits for January 13, 2007. - 45 -

Figure 14a. - 46 -

Figure 14b. (Top) Waveform fits for 13 January 2007 for the preferred slip model (Bottom). Data are shown by black lines and model predictions by red lines. Teleseismic P and SH waves resolve shorter scale structure of the source slip function, while dispersion-corrected Rayleigh pulses constrain overall duration and centroid properties of the solution. The source moment rate function is displayed above the slip model with time scale corresponding to the isochrones of the rupture front that spread at 3.5 km/s. - 47 -

Figure 15. (a) Moment rate spectra for the 15 November 2006 and 13 January 2007 events. Note the larger high frequency amplitudes for the smaller January event. This is associated with higher energy release and higher energy/seismic moment ratio for the January event. (b) Source time functions for the doublet events. Note the differences in total duration and overall complexity. - 48 -

Figure 16. Surface-projection of coseismic slip for the 15 November 2006 (average slip 4.6 m) and 13 January 2007 (average slip 9.6 m) events (NEIC epicenters: yellow circles, CMT centroid epicenters: stars). CMT mechanisms (centered on NEIC epicenters) for large events between June 2006 and May 2007 are shown; enlarged versions for the doublet events. Gray mechanisms indicate events before the 15 November 2006 event, red mechanisms indicate events after that rupture. The focal mechanism and epicenter of the 16 March 1963 compressional outer-rise event (hexagon) are included. The arrow indicates the direction of the Pacific plate motion at 80 mm/yr. - 49 -

Figure 17. Summary cartoon showing regional seismicity and structural interpretation of the 2006-2007 Kuril Islands earthquake sequence. - 50 -