geosphere subduction zone megathrust earthquakesthorne/tl.pdfs/bl... · 2019-07-24 · research...

33
Research Paper 1468 Bilek and Lay | Subduction zone megathrust earthquakes GEOSPHERE | Volume 14 | Number 4 Subduction zone megathrust earthquakes Susan L. Bilek 1 and Thorne Lay 2 1 Department of Earth and Environmental Science, New Mexico Tech, Socorro, New Mexico 87801, USA 2 Department of Earth and Planetary Sciences, University of California–Santa Cruz, Santa Cruz, California 95064, USA ABSTRACT Subduction zone megathrust faults host Earth’s largest earthquakes, along with multitudes of smaller events that contribute to plate convergence. An understanding of the faulting behavior of megathrusts is central to seismic and tsunami hazard assessment around subduction zone margins. Cumula- tive sliding displacement across each megathrust, which extends from the trench to the downdip transition to interplate ductile deformation, is accom- modated by a combination of rapid stick-slip earthquakes, episodic slow-slip events, and quasi-static creep. Megathrust faults have heterogeneous fric- tional properties that contribute to earthquake diversity, which is considered here in terms of regional variations in maximum recorded magnitudes, Guten- berg-Richter b values, earthquake productivity, and cumulative seismic mo- ment depth distributions for the major subduction zones. Great earthquakes on megathrusts occur in irregular cycles of interseismic strain accumulation, foreshock activity, main-shock rupture, postseismic slip, viscoelastic relax- ation, and fault healing, with all stages now being captured by geophysical monitoring. Observations of depth-dependent radiation characteristics, large earthquake slip distributions, variations in rupture velocities, radiated energy and stress drop, and relationships to aftershock distributions and afterslip are discussed. Seismic sequences for very large events have some degree of regularity within subduction zone segments, but this can be complicated by supercycles of intermittent huge ruptures that traverse segment boundaries. Factors influencing variability of large megathrust ruptures, such as large-scale plate structure and kinematics, presence of sediments and fluids, lower-plate bathymetric roughness, and upper-plate structure, are discussed. The diver- sity of megathrust failure processes presents a suite of natural hazards, in- cluding earthquake shaking, submarine slumping, and tsunami generation. Improved monitoring of the offshore environment is needed to better quantify and mitigate the threats posed by megathrust earthquakes globally. INTRODUCTION The contact surfaces between underthrusting and overriding plates in subduction zones are called megathrust faults (Fig. 1). Over time, aside from secondary inelastic deformation or block motions of the surrounding plates, the total displacement across the entire megathrust surface must equal the relative plate convergence. This slip budget extends over the depth range from the subduction zone trench to the downdip limit of the sliding plate boundary, defined by the onset of ductile flow in the mantle wedge. The megathrust itself may involve a shear zone rather than a single surface, and repeated ruptures of the plate-boundary contact may occur on parallel faults within the shear zone. The overall plate convergence is accommodated by spatially varying stick-slip earthquake failure, episodic slow slip, and continuous aseismic slid- ing. Heterogeneous frictional properties of the megathrust and dynamic inter- actions control how any given region behaves, including the possibility that some regions of stable sliding might transition to stick-slip failure during high- strain-rate loading. Large-scale deformation of the underthrusting plate occurs in subduction zones, with the plate bending and straightening as it subducts beneath either oceanic or continental lithosphere, such that the forces acting on each megathrust and the resulting seismicity are affected by the broad tec- tonic configuration of each plate boundary. All of these deformation processes can produce hazards, from seismic shaking to submarine slumping and tsu- nami generation. Here, we focus on megathrust faulting, which produces the greatest hazards. Based on the elastic rebound theory of earthquake occurrence and the rel- atively steady loading provided by long-term plate convergence in subduction zones, one might expect the stick-slip failure of megathrust faults to be at least somewhat regular. This does appear to hold for many regions, where irregular “cycles” of long intervals of interseismic strain accumulation, sudden earth- quake sliding, and postseismic adjustment are apparent for well-documented sequences of great earthquakes (M > 8) in regions such as the Nankai Trough of southwest Japan, along southern Chile, and along the Kuril Islands (e.g., Ando, 1975; McCann et al., 1979; Lay et al., 1982). In many places, the historical record of great earthquakes is limited, with too few events to establish any reg- GEOSPHERE GEOSPHERE; v. 14, no. 4 https://doi.org/10.1130/GES01608.1 19 figures; 2 tables CORRESPONDENCE: [email protected] CITATION: Bilek, S.L., and Lay, T., 2018, Subduction zone megathrust earthquakes: Geosphere, v. 14, no. 4, p. 1468–1500, https://doi.org/10.1130/GES01608.1. Science Editor: Shanaka de Silva Guest Associate Editor: Laura M. Wallace Received 26 August 2017 Revision received 19 April 2018 Accepted 19 June 2018 Published online 6 July 2018 OPEN ACCESS GO L D This paper is published under the terms of the CC‑BY‑NC license. © 2018 The Authors 0 50 100 150 200 25 50 Trench upper plate mantle crust crust seismogenic zone Depth (km) Distance from Trench (km) shallow slip tsunami earthquakes slow slip large/great earthquakes slow slip tremor Figure 1. Megathrust fault zone within a generalized subduction zone, highlighting the diverse slip modes observed in the shallow seismogenic, or earthquake-producing, region. THEMED ISSUE: Subduction Top to Bottom 2

Upload: others

Post on 24-Jun-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1468Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

Subduction zone megathrust earthquakesSusan L. Bilek1 and Thorne Lay2

1Department of Earth and Environmental Science, New Mexico Tech, Socorro, New Mexico 87801, USA2Department of Earth and Planetary Sciences, University of California–Santa Cruz, Santa Cruz, California 95064, USA

ABSTRACT

Subduction zone megathrust faults host Earth’s largest earthquakes, along with multitudes of smaller events that contribute to plate convergence. An understanding of the faulting behavior of megathrusts is central to seismic and tsunami hazard assessment around subduction zone margins. Cumula-tive sliding displacement across each megathrust, which extends from the trench to the downdip transition to interplate ductile deformation, is accom-modated by a combination of rapid stick-slip earthquakes, episodic slow-slip events, and quasi-static creep. Megathrust faults have heterogeneous fric-tional properties that contribute to earthquake diversity, which is considered here in terms of regional variations in maximum recorded magnitudes, Guten-berg-Richter b values, earthquake productivity, and cumulative seismic mo-ment depth distributions for the major subduction zones. Great earthquakes on megathrusts occur in irregular cycles of interseismic strain accumulation, foreshock activity, main-shock rupture, postseismic slip, viscoelastic relax-ation, and fault healing, with all stages now being captured by geophysical monitoring. Observations of depth-dependent radiation characteristics, large earthquake slip distributions, variations in rupture velocities, radiated energy and stress drop, and relationships to aftershock distributions and afterslip are discussed. Seismic sequences for very large events have some degree of regularity within subduction zone segments, but this can be complicated by supercycles of intermittent huge ruptures that traverse segment boundaries. Factors influencing variability of large megathrust ruptures, such as large-scale plate structure and kinematics, presence of sediments and fluids, lower-plate bathymetric roughness, and upper-plate structure, are discussed. The diver-sity of megathrust failure processes presents a suite of natural hazards, in-cluding earthquake shaking, submarine slumping, and tsunami generation. Improved monitoring of the offshore environment is needed to better quantify and mitigate the threats posed by megathrust earthquakes globally.

INTRODUCTION

The contact surfaces between underthrusting and overriding plates in subduction zones are called megathrust faults (Fig. 1). Over time, aside from secondary inelastic deformation or block motions of the surrounding plates, the total displacement across the entire megathrust surface must equal the relative plate convergence. This slip budget extends over the depth range from

the subduction zone trench to the downdip limit of the sliding plate boundary, defined by the onset of ductile flow in the mantle wedge. The megathrust itself may involve a shear zone rather than a single surface, and repeated ruptures of the plate-boundary contact may occur on parallel faults within the shear zone. The overall plate convergence is accommodated by spatially varying stick-slip earthquake failure, episodic slow slip, and continuous aseismic slid-ing. Heterogeneous frictional properties of the megathrust and dynamic inter-actions control how any given region behaves, including the possibility that some regions of stable sliding might transition to stick-slip failure during high-strain-rate loading. Large-scale deformation of the underthrusting plate occurs in subduction zones, with the plate bending and straightening as it subducts beneath either oceanic or continental lithosphere, such that the forces acting on each megathrust and the resulting seismicity are affected by the broad tec-tonic configuration of each plate boundary. All of these deformation processes can produce hazards, from seismic shaking to submarine slumping and tsu-nami generation. Here, we focus on megathrust faulting, which produces the greatest hazards.

Based on the elastic rebound theory of earthquake occurrence and the rel-atively steady loading provided by long-term plate convergence in subduction zones, one might expect the stick-slip failure of megathrust faults to be at least somewhat regular. This does appear to hold for many regions, where irregular “cycles” of long intervals of interseismic strain accumulation, sudden earth-quake sliding, and postseismic adjustment are apparent for well-documented sequences of great earthquakes (M > 8) in regions such as the Nankai Trough of southwest Japan, along southern Chile, and along the Kuril Islands (e.g., Ando, 1975; McCann et al., 1979; Lay et al., 1982). In many places, the historical record of great earthquakes is limited, with too few events to establish any reg-

GEOSPHERE

GEOSPHERE; v. 14, no. 4

https://doi.org/10.1130/GES01608.1

19 figures; 2 tables

CORRESPONDENCE: Susan .Bilek@ nmt .edu

CITATION: Bilek, S.L., and Lay, T., 2018, Subduction zone megathrust earthquakes: Geosphere, v. 14, no. 4, p. 1468–1500, https:// doi .org /10 .1130 /GES01608.1.

Science Editor: Shanaka de SilvaGuest Associate Editor: Laura M. Wallace

Received 26 August 2017Revision received 19 April 2018Accepted 19 June 2018Published online 6 July 2018

OPEN ACCESS

GOLD

This paper is published under the terms of the CC‑BY‑NC license.

© 2018 The Authors

0 50 100 150 200

25

50

Trench

upper plate

mantlecrust

crust seismogenic zone

Dep

th (k

m)

Distance from Trench (km)

shallow sliptsunami earthquakesslow slip

large/great earthquakes

slow sliptremor

Figure 1. Megathrust fault zone within a generalized subduction zone, highlighting the diverse slip modes observed in the shallow seismogenic, or earthquake-producing, region.

THEMED ISSUE: Subduction Top to Bottom 2

Page 2: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1469Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

ularity, or such limited information about early events that it is hard to establish their relationship to more recent well-documented ruptures. Nonetheless, the time interval since the last great earthquake in a given region and that event’s inferred size and spatial extent have quite successfully guided designation of “seismic gaps” as regions with potential for future great earthquakes (e.g., McCann et al., 1979; Nishenko, 1991), albeit without time predictability (e.g., Lay, 2015). Smaller events appear to be irregular in occurrence and to interact strongly with nearby ruptures, causing the seismic gap approach to be less useful than for great events. This perspective is not lacking in contention, as some researchers deny any degree of cyclicity in earthquake occurrence (e.g., Kagan and Jackson, 1995; Kagan et al., 2012), but for very large earthquakes, these objections do not appear to be valid. Regularity in great earthquake cy-cles has long been recognized to be complicated by intermittent triggering of adjacent megathrust segments that otherwise tend to fail separately in smaller events, notably in southwest Japan (Ando, 1975) and Ecuador-Colombia (Kanamori and McNally, 1982), with synchronization of rupture possibly oc-curring in supercycles of huge earthquakes on time scales that greatly exceed recorded history (e.g., Sieh et al., 2008). Such unprecedented events have low, but nonzero probability for many regions, and it is difficult to reliably evaluate the worst-case potential and its associated probability of occurrence, given the limited temporal span of the seismological and paleoseismological records.

Surprises that stem from the limited historic record for megathrust events have included great earthquakes striking in regions lacking any known prior great event and having disrupted subduction zone structure. This includes examples such as subduction of a young ridge at a triple junction where a great earthquake struck in the Solomon Islands in 2007 (e.g., Taylor et al., 2008; Furlong et al., 2009), and the great 2006 Kuril Islands rupture (e.g., Ammon et al., 2008), which struck in a region with discontinuous arc structure and an unusual forearc basin. Both areas had been argued to potentially be perma-nently aseismic, an assertion also still invoked for regions of subducting ridges such as the Tehuantepec and Carnegie Ridges along South America, where there is no history of great earthquakes. Confidence in such assertions must be assessed with caution, and the use of geodetic measurements to assess upper-plate strain is the most promising approach to evaluating actual seismic potential in such regions.

Great megathrust earthquakes have also now been observed to have trig-gering interactions with great intraplate faulting (e.g., Ammon et al., 2008; Lay et al., 2010b, 2017), and cascading failures on the relevant megathrust (e.g., Lay et al., 2010a), presenting further challenges to hazard assessment and rapid warning procedures. Very large earthquakes have also been documented to rupture the shallowest part of the megathrust (e.g., Polet and Kanamori, 2000; Lay and Bilek, 2007), where it had been thought that aseismic deformation of wet sediments on the plate boundary would preclude stick-slip sliding.

Deployment of high-quality geodetic networks in subduction zones has added important constraints on coseismic slip during nearby megathrust earthquakes, along with capturing deformation prior to and after large rup-tures of the plate boundary. This has revealed the existence of slow-slip events

that may be accompanied by seismic tremor or small repeating earthquakes, but the full deformation occurring over days to months can only be resolved geodetically. Deformation in the upper plate of a frictionally locked megathrust, with landward compression, can be resolved by global positioning system (GPS) and Global Navigation Satellite System (GNSS) sensors operating over years to decades, and these data sets now provide direct measurement of the interseismic process and the ability to establish whether a seismic gap is actu-ally accumulating strain that will result in a future large earthquake or whether aseismic sliding is releasing the strain. Geodesy also enables characterization of upper-plate sliver or block motions that may result from oblique conver-gence. This is now recognized as quite common, and it must be accounted for when evaluating the megathrust strain energy budget (e.g., McCaffrey et al., 2000; Wallace et al., 2004; La Femina et al., 2009). Precursory sliding near a fu-ture hypocenter, postseismic slip around main-shock rupture zones, relocking of a ruptured zone, and viscoelastic deformation after an event are processes that have been sensed by onshore and, increasingly, offshore geodetic mea-surements, greatly expanding our view into the seismic strain accumulation and release process and the ways in which relative plate motion is accommo-dated across the entire megathrust (e.g., Ito et al., 2013; Yokota et al., 2016). In this overview, we highlight some of the basic seismological characteristics of major subduction zones and salient findings of recent investigations of large earthquakes and slow deformation on megathrusts. We conclude with consid-eration of the seismic hazards posed by megathrust earthquakes.

SEISMICITY CHARACTERISTICS OF MEGATHRUST FAULTS

Earthquake occurrence reveals many attributes of subduction zone megathrusts, although it is good to keep in mind that cumulative seismic slip appears to account for less than half of total convergence across the global set of plate-boundary faults in the seismic record. Variations in earthquake size, number, and depth distribution indicate distinct properties of megathrust faults that guide comparisons of processes in different subduction zones.

Largest-Magnitude Ruptures

Measured seismic magnitudes for shallow subduction zone earthquakes span from near zero at the low end, depending on availability of regional seis-mic monitoring, to the largest observed event, the 1960 Chile moment magni-tude (Mw) = 9.5 earthquake. Globally, earthquake catalogs have improved with time, with fairly complete catalogs back to 1900 for events larger than M ~ 7, and near-complete recording since the mid-1970s of events larger than M ~ 5.0. To provide a sampling of diverse megathrust activity, we considered the seismicity characteristics for nine major subduction zones (Fig. 2; Table 1) and identified the largest observed earthquakes in each zone using two readily available earth-quake catalogs, the U.S. Geological Survey National Earthquake Information

Page 3: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1470Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

1964

90°

90°

120°

120°

150°

150°

180°

180°

–150° –120°

–150° –120°

–90° –60°

–90° –60°

−60° −60°

−30° −30°

0° 0°

30° 30°

60° 60°

1923

1938

1946 1957

1986

1965

Aleutian-Alaska

Chile

Peru

Mexico-Middle America

Tonga-Kermadec

Sumatra

Honshu

Kuril-Kamchatka

2012

2004 2005

2007

Marianas

1976

1906

1985 1995

2001

2015, 1943 2010

1960 (2)

1906

1995 2014

1985

1917 2009

1922

2007

1917

1919

1932

1942

2011

1958 1963

2006 2007

1944, 1946

19401966

193319681958, 2003

1952

1960

1923

1994

2013 2007

1918 Figure 2. Great (Mw ≥ 8.0) earthquakes from 1900 to 2016 presumed to be on subduc-tion zone megathrusts (green circles) and some important intraplate ruptures (red circles) discussed in the text, from the U.S. Geological Survey National Earthquake Information Center catalog (USGS-NEIC) catalog. Events mentioned in the text are labeled with dates. Some events in the Southwest Pacific subduction zones have uncertain faulting parameters, but most are likely thrust events in the small subduc-tion zones in the region. No great events in the Marianas subduction zone have been recorded. Note that regions such as Cascadia, New Zealand, Izu-Bonin, Luzon, Central America, and the Himalayan front are not represented because great events have not happened in these regions in the seismological time period of record. Histori-cal great ruptures have been documented in Cascadia and the Himalayas.

TABLE 1. MAXIMUM AND MEDIAN EARTHQUAKE MAGNITUDES FROM THE GLOBAL CENTROID MOMENT TENSOR (GCMT) CATALOG

RegionLatitude range

(°)Longitude range

(°) Trench strikeMax. Mw

(all)Max. Mw

(megathrust)Median Mw

(all)Median Mw

(megathrust)

Alaska 49.17N–61.65N 165.06E–149.11W 250 ± 40° 8.0 8.0 5.3 5.4Central America 12N–20N;

7.6N–12N106.12W–94.44W;

83.47W–94.44W290 ± 30° 8.0 8.0 5.4 5.4

Chile 18S–48S 66W–78W 0 ± 30° 8.8 8.8 5.5 5.5Japan 34.5N–41N 140E–145E 190 ± 30° 9.1 9.1 5.3 5.3Kuril/Kamchatka 41N–55.67N 141.47E–163.70E 210 ± 40° 8.3 8.3 5.3 5.3Marianas 10.45N–25.10N 138.17E–150.5E 240 ± 30° (10–13N);

160 ± 40° (13N–22N);145 ± 30° (22N–25.1N)

7.7 7.1 5.3 5.3

Peru 2.2S–18.5S 82.78W–67.25W 330 ± 30° 8.4 8.4 5.3 5.4Sumatra 5S–15N 90E–108E 320 ± 30° 9.0 9.0 5.2 5.3Tonga-Kermadec 14.3S–35.3S 176.6E–170.5W 200 ± 30° 8.1 8.0 5.3 5.4

Note: Depth range is 0–100 km, and time window is 1 January 1976 to 7 December 2016 for all regions. The subset designated megathrust is for thrust events with focal mechanism consistent with local trench strike determined from bathymetric data. Mw is moment magnitude.

Page 4: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1471Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

Center catalog (USGS-NEIC, https:// earthquake .usgs .gov), from 1900 to 2016, and the Global Centroid Moment Tensor catalog (GCMT, http:// www .globalcmt .org), from 1976 to 2016. The USGS-NEIC catalog includes the largest- magnitude earthquakes over the last century, drawing upon redetermined magnitude and location estimates from the International Seismological Centre–Global Earth-quake Model (ISC-GEM; http:// www .isc .ac .uk /iscgem/). The GCMT catalog pro-vides relatively uniformly determined estimates of each event’s moment tensor and seismic moment. We utilized the focal mechanisms to identify events that were most likely located on the subduction megathrust (Figs. 3–11).

Great Earthquake Catalog (1900–2016)

We first considered instrumentally recorded, great (M ≥ 8.0), shallow plate-boundary thrust-faulting events in nine major subduction zones (Fig. 2; Table 1). For earthquake seismologists, most of these earthquakes are immedi-ately identifiable by their years and place names, as they represent major tec-tonic events that have sometimes produced devastating shaking and tsunami damage. Along the Alaska-Aleutian subduction zone, six great earthquakes spanned the arc in 1964 (Alaska, Mw 9.2), 1965 (Rat Islands, Mw 8.7), 1957 (Fox

165° 170° 175° 180° –175° –170° –165° –160° –155° –150°

50°

55°

60°

Alln = 1425b = 0.82

Thrustn = 893b = 0.79

log(

#Eve

nts/

year

Mag

nitu

de >

Mw

)

Mw

2.0

1.5

1.0

0.5

0.0

–0.5

–1.0

–1.54.0 4.5 5.0 5.5 6.0 6.5 7.0 7.5 8.0 8.5

Mw

–1.54.0 4.5 5.0 5.5 6.0 6.5 7.0 7.5 8.0 8.5

1.5

1.0

0.5

0.0

–0.5

–1.0

100 km

Figure 3. Thrust faulting distribution and Mw-frequency statistics for the Alaska- Aleutian subduction zone from the 1976–2016 Global Centroid Moment Tensor (GCMT) catalog. Top: Map of earthquakes that have thrust mechanisms (lower-hemi-sphere best double-couple mechanisms), most of which are consistent with the lo-cal trench orientation, along with all earth-quakes in the catalog for this region (small red circles). Bottom: Earthquake cumula-tive magnitude distributions for all GCMT earthquakes in the region (left) and for the thrust earthquakes shown in above map (right); the number of earthquakes in each set is n, and b is the Gutenberg- Richter b value estimated by the maximum like-lihood method discussed in the text. The GCMT catalog is considered complete for Mw ≥ 5.2, and we used this data set for the b value calculation (dark-blue circles).

Page 5: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1472Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

Islands, Mw 8.6), 1946 (Mw 8.6), 1938 (Alaskan Peninsula, M 8.3), and 1986 (Fox Islands, Mw 8.0). The 1986 event reruptured a western portion of the much larger 1957 earthquake. The 1946 event ruptured between the 1957 and 1938 events and produced a disproportionally large tsunami, leading to it being identified as a “tsunami earthquake” by Kanamori (1972). These events have essentially filled in most of the Alaska-Aleutian plate boundary during this cen-tury. An historical great event in 1788 ruptured the 1938 zone and Kodiak Island portions of the 1964 ruptures (e.g., Briggs et al., 2014), indicating that bound-aries between recent ruptures have not been persistent.

Along Mexico and Middle America, the largest seismically recorded earth-quakes have been much smaller, and they occurred along the coast of Mex-ico—an M 8.1 earthquake in 1932 (Jalisco) and two Mw 8 earthquakes in 1985 (Michoacan) and 1995 (Colima). However, the largest documented historical event occurred along the coast from Guerrero to Tehuantepec in 1787 with M ~ 8.6 (Suárez and Albini, 2009), apparently rupturing across multiple zones of more recent smaller events, so the limited time span of the seismologi-cal catalog must be kept in mind. The largest recorded earthquake in Peru was the 2001 (Arequipa, Mw 8.4) event, which ruptured within approximately

–106° –104° –102° –100° –98° –96° –94° –92° –90° –88° –86° –84°

10°

12°

14°

16°

18°

20°

Thrustn = 659b = 0.71

Alln = 1242b = 0.76

log(

#Eve

nts/

year

Mag

nitu

de >

Mw

)

Mw

2.0

1.5

1.0

0.5

0.0

–0.5

–1.0

–1.5

4.5 5.0 5.5 6.0 6.5 7.0 7.5 8.0 8.5–2.0

Mw4.5 5.0 5.5 6.0 6.5 7.0 7.5 8.0 8.5

1.5

1.0

0.5

0.0

–0.5

–1.0

–1.5

–2.0

100 km

Figure 4. Thrust faulting earthquake distri-bution and Mw-frequency distributions for the Mexico–Middle America subduction zone from the Global Centroid Moment Tensor (GCMT) catalog. See description of Figure 3 for details.

Page 6: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1473Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

two-thirds of the zone of the 1868 (Arica, M ~8.5–9.0) earthquake. Other great Peruvian earthquakes occurred in 1940 (Mw 8.2), 1966 (Mw 8.1), 1942 (Mw 8.1), and 2007 (Pisco, Mw 8.0). The Chile subduction zone has experienced 10 great events: 1960 (Valdivia, Mw 9.5), 2010 (Maule, Mw 8.8), 1922 (Atacama, M 8.5), 2015 ( Illapel, Mw 8.3), 1906 (Valparaíso, M 8.2), 2014 (Iquique, Mw 8.2), 1960 (Bio-Bio, Mw 8.1), 1943 (Coquimbo, M 8.1), 1985 (Valparaíso, Mw 8.0), and 1995 (Antofagasta, Mw 8.0). Of these, only the 1943 and 2015 events rupture zones overlapped significantly. The region south of the 2014 Iquique event last rup-tured in 1877 (Iquique, M ~8.8), and this remains a major seismic gap. The 1922 rupture zone along Atacama likely has substantial strain accumulation relative to surrounding regions, as geodetic measurements indicate that the mega-thrust is pervasively locked in the region (Métois et al., 2014).

Along the Tonga-Kermadec arc, the largest known interplate thrust event occurred in 1976 (Kermadec, Mw 8.0). The 2009 Samoa earthquake (Mw 8.1)

began with interplate normal faulting but triggered Mw 8.0 thrusting on the northern Tonga megathrust (Beavan et al., 2010; Lay et al., 2010b; Fan et al., 2016). Other great shallow events in the region in 1917 (Kermadec, M 8.2), 1919 (Tonga, M 8.1), and 1917 (Samoa, M 8.0) all appear to have been intraplate events (Meng et al., 2015b). The subduction zones along Vanuatu, Solomon Is-lands, Sulawesi, Taiwan, and the Philippines all have maximum recorded shal-low thrusting earthquake magnitudes of about M ~8.1. Maximum magnitudes for recorded events along the Marianas subduction zone have been less than M 8. While it is conceivable that very-low-probability events might rupture the entire megathrust contact in each of these island-arc regions, it may be that the apparently low coupling and high fraction of aseismic convergence preclude this from happening, and seismic hazard assessments can focus on ruptures of the size historically observed. In contrast, the Sumatra subduction zone has experienced several great thrust earthquakes since 2004, in 2004 (Sumatra-

–78°–76°–74°–72°–70°–68°–66°–44°

–42°

–40°

–38°

–36°

–34°

–32°

–30°

–28°

–26°

–24°

–22°

–20°

–18°All

n = 1175b = 0.77

Thrustn = 721b = 0.73

log(

#Eve

nts/

year

Mag

nitu

de >

Mw

)

Mw

2.0

1.5

1.0

0.5

0.0

–0.5

–1.0

–1.5

4.5 5.0 5.5 6.0 6.5 7.0 7.5 8.0 8.5–2.0

log(

#Eve

nts/

year

Mag

nitu

de >

Mw

)

Mw

1.5

1.0

0.5

0.0

–0.5

–1.0

–1.5

4.5 5.0 5.5 6.0 6.5 7.0 7.5 8.0 8.5–2.0

9.0

9.0

100 km

Figure 5. Thrust-faulting earthquake distri-bution and Mw-frequency distributions for the Chile subduction zone from the Global Centroid Moment Tensor (GCMT) catalog. See description of Figure 3 for details.

Page 7: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1474Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

Andaman, Mw 9.15), 2005 (Nias, Mw 8.6), and 2007 (Bengkulu, Mw 8.4). Multiple historical great thrust earthquakes of up to M ~9 struck the region prior to 1900 as well. Great intraplate strike-slip ruptures with Mw 8.6 and 8.2 struck in 2012 in the Indo-Australian deformation zone in the oceanic plate south of Sumatra.

The Honshu, Japan, subduction zone has hosted three great thrust earth-quakes since 1900, with the largest being the 2011 (Tohoku, Mw 9.1) event in 2011, followed by the 1968 (Tokachi-oki, Mw 8.2) and 1960 (Sanriku, Mw 8.0; possibly an overestimated magnitude) earthquakes. The 2011 event appears to have reruptured a significant portion of the 1896 Meiji Sanriku tsunami earth-quake zone on the shallow portion of the megathrust near the trench from 38°N to 40.3°N (e.g., Yamazaki et al., 2018; Lay, 2017). Other significant north-ern Japan earthquakes occurred in 1933 (Sanriku outer-rise normal faulting, M 8.4) and 1923 (Kanto oblique thrust along the Sagami Trough, M 8.1). Nine great thrust earthquakes have occurred along the Hokkaido-Kuril-Kamchatka

subduction zone, including an Mw 9.0 event in 1952 (Kamchatka), and the 1963 (Kuril Islands, Mw 8.5), 1923 (Kamchatka, M 8.4), 2006 (Central Kuril Islands, Mw 8.3), 2003 (Hokkaido, Mw 8.3), 1994 (Kuril Islands Mw 8.3), 1958 (Kuril Islands, Mw 8.3), 1952 (Hokkaido, Mw 8.1), and 1918 (Central Kuril Islands, M 8.1) events. The 2006 Kuril Islands event appears to have triggered the 2007 (Mw 8.1) outer- rise normal fault rupture (Ammon et al., 2008; Lay et al., 2009), given that it im-mediately activated outer-rise faulting aftershocks. The 2003 Hokkaido rupture zone closely overlaps the 1952 zone.

The variation in largest recorded earthquake size between regions is very substantial, and many efforts have been made to establish the factors influenc-ing the distribution of great earthquakes, as discussed later. Within the seis-mological record, only a few recent great events have reruptured a region with seismological recordings of prior great earthquake activity (1986 Aleutians, 2015 Illapel, 2003 Hokkaido). Because the prior events (1957, 1943, 1952, re-

140° 142° 144°

36°

38°

40°

Thrustn = 720b = 0.75

Alln = 1257b = 0.81

log(

#Eve

nts/

year

Mag

nitu

de >

Mw

)

Mw

2.0

1.5

1.0

0.5

0.0

–0.5

–1.0

–1.5

4.0 5.0 6.0 7.0 8.0–2.0

log(

#Eve

nts/

year

Mag

nitu

de >

Mw

)

1.5

1.0

0.5

0.0

–0.5

–1.0

–1.5

–2.0

9.0

Mw4.0 5.0 6.0 7.0 8.0 9.0

100 km

Figure 6. Thrust faulting earthquake distribution and Mw-frequency distribu-tions for the Honshu, Japan, subduction zone from the Global Centroid Moment Tensor (GCMT) catalog. See description of Figure 3 for details.

Page 8: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1475Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

spectively) all have limited seismological data and relatively poor constraints on their rupture characteristics, it is very difficult to directly evaluate the sta-bility of repeated great ruptures of a given portion of the relevant megathrust.

It is important to recognize that only ~118 yr of reliable seismological rec-ords exist, and many earlier great events have been documented around the Pacific subduction zones by historical accounts and paleoseismic analyses (e.g., Beck et al., 1998; Kelleher, 1972; Lay et al., 1982; McCann et al., 1979; Scholz and Campos, 1995, 2012). In some cases, such as for the 1837 and 1737 events in the 1960 Chile rupture zone, the size of the historic events is being reevaluated based on detailed examination of the record, modifying inferences of segmentation and regularity (e.g., Cisternas et al., 2017). The Cascadia subduction zone is known to have had many great earthquakes, but the most recent was in 1700, and so it is not represented by the seismological catalogs. Similarly, the Himalayan front hosts great thrust events, but none has occurred during the seismological record. Southwest Japan did have well-characterized great earthquakes in 1944 and 1946, and documentation and field observations exist for numerous events in the same region dating back over 1000 yr (e.g., Ando, 1975).

GCMT Catalog (1976–2016)

During the relatively short 40 yr time interval spanned by the GCMT cata-log, the largest-magnitude megathrust earthquakes occurred in Sumatra (2004 Mw 9.2), northern Japan (2011 Mw 9.1), and central Chile (2010 Mw 8.8). In Fig-ures 3–11, we show GCMT faulting mechanisms for likely megathrust events (those having shallow-dipping thrust fault solutions) for our subset of nine subduction zones, along with regional Mw-frequency distributions for shallow GCMT events. The maps display the variation in megathrust geometry and width among the major subduction zones. Table 1 lists details of the regions and magnitude distributions from this catalog.

b Values

Frequency-magnitude distributions are often characterized by the slope of the linear fit to the logarithmic number of earthquakes equal to or greater than a given magnitude (Gutenberg and Richter, 1944). This slope, the b value, is typically around 1 for global and regional earthquake catalogs, although there

142° 144° 146° 148° 150° 152° 154° 156° 158° 160° 162°

42°

44°

46°

48°

50°

52°

54°

Thrustn = 1203b = 0.80

Alln = 1666b = 0.83

log(

#Eve

nts/

year

Mag

nitu

de >

Mw

)

Mw

2.0

1.5

1.0

0.5

0.0

–0.5

–1.0

–1.5

4.0 5.0 6.0 7.0 8.0–2.0

log(

#Eve

nts/

year

Mag

nitu

de >

Mw

) 1.5

1.0

0.5

0.0

–0.5

–1.0

–1.5

–2.0

4.5 5.5 6.5 7.5 8.5

Mw4.0 5.0 6.0 7.0 8.04.5 5.5 6.5 7.5 8.5

100 km

Figure 7. Thrust faulting earthquake distri-bution and Mw-frequency distributions for the Hokkaido-Kuril-Kamchatka subduction zone from the Global Centroid Moment Tensor (GCMT) catalog. See description of Figure 3 for details.

Page 9: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1476Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

can be some variation. Variability in b value has been used to assess variations in stress conditions on the fault (e.g., Frohlich and Davis, 1993; Schorlemmer and Wiemer, 2005; Schorlemmer et al., 2005; Ghosh et al., 2008). We used the GCMT catalog centroid locations, magnitudes, and focal mechanisms for all events with depths <100 km. We computed b values for all GCMT events in each region, as well as for all thrust events, most of which are likely to have oc-curred on the subduction megathrusts based on location and faulting mecha-nism for each region (Figs. 3–11). The mechanism selection criteria for designa-tion as “likely located on the megathrust” are a 45° to 90° plunge of the tension axis, and a strike within 30° of the regional trench strike. Using the data sets of all events or only thrust events, we computed b values using the  Aki-Utsu equation (Aki, 1965; Utsu, 1965):

be

M MdM

c

=− −

log10

2

, (1)

where M is the mean magnitude of the regional distribution for events above a completeness threshold Mc, and dM  = 0.1 (controlled by rounding of the magnitudes; Table 2). We set Mc = 5.2 for both populations, based on prior es-

timates of the GCMT catalog completeness (Ekström et al., 2012; Wetzler et al., 2017). This choice appears to be adequate for all of the cases shown here, and use of a higher Mc value did not change the relative values significantly.

In all regions except for the Marianas, we found b values < 1 for both the total and megathrust populations (Table 2). The numbers of events in the pop-ulations and the magnitude ranges spanned by the data sets (see Figs. 3–11) are limited for the GCMT catalog, and most magnitude distributions only have well-sampled distributions up to Mw ~ 7.5 in the figures (lower in the Marianas). These data limitations give rise to significant uncertainty in the b values of ~0.05, so regional differences are suggested, but not well resolved. Apart from the Marianas and Tonga, the b values for the megathrust events (0.7–0.8) were found to be systematically lower than for the full populations (0.76–0.88) by an average of 0.07. The Tonga subduction zone has b values of 0.92 and 0.93 for the full data set and the megathrust, respectively, while the sparse Marianas data set has corresponding b values of 1.03 and 1.07, respectively. Reduction of the b value for the quasi-two-dimensional, single-fault-plane thrust faulting data set relative to the three-dimensional total distribution is plausibly the re-sult of having more faulting geometries in the greater volumes sampled by the full data set, although the differences are marginally resolved. The high thrust-

140° 142° 144° 146° 148° 150°

12°

14°

16°

18°

20°

22°

24°

Thrustn = 181b = 1.07

Alln = 732b = 1.03

log(

#Eve

nts/

year

Mag

nitu

de >

Mw

) 1.5

1.0

0.5

0.0

–0.5

–1.0

–1.5

–2.0

Mw5.0 6.0 7.0 8.04.5 5.5 6.5 7.5

log(

#Eve

nts/

year

Mag

nitu

de >

Mw

) 1.0

0.5

0.0

–0.5

–1.0

–1.5

–2.0

–2.5

Mw5.0 6.0 7.04.5 5.5 6.5 7.5

100 km

Figure 8. Thrust faulting earthquake distri-bution and Mw-frequency distributions for the Marianas subduction zone from the Global Centroid Moment Tensor (GCMT) catalog. See description of Figure 3 for details.

Page 10: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1477Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

fault b values for the Tonga (Fig. 11) and Marianas (Fig. 8) zones, which have steeply dipping and narrow megathrust interfaces, are likely due to inclusion of many thrust faulting intraplate events by our criteria, since intraslab com-pression events are known to occur in these regions (e.g., Meng et al., 2015a).

Previous studies of b values in subduction zones have also found b < 1 in shallow megathrust zones. Along the northern Japan subduction zone, Tor-mann et al. (2015) estimate b values of 0.5–0.9 at depths between the surface and ~100 km, although they did not specifically separate out events occurring on the megathrust zone versus upper-plate or intraslab events. Nishikawa and Ide (2014) computed b values for all of the subduction zones considered here, although with more spatial subdivisions within each subduction zone and no separation by mechanism. Given the data and methodological differences in b value calculation, comparison of absolute values is difficult. However, their lowest b values (~0.8 for parts of Central America and Peru) correspond to relatively low b values in our analysis, and their highest b values, exceeding 1, were also found for the Marianas and Tonga-Kermadec regions.

Regional Seismic Productivity

We also used the GCMT Mw-frequency relations for Mc = 5.2 to evaluate overall seismic productivity variations during the 1976–2016 time period using both the full set of events, as well as only thrust faulting events. The Japan and Tonga-Kermadec subduction zones are most productive overall, with ~1.5 earthquakes of any mechanism per kilometer length along strike of the subduc-tion zone. Sumatra is next, with ~0.7 events/km, followed by Kuril-Kamchatka (0.6 events/km), Chile (0.5 events/km), Mexico–Middle America (0.4 events/km), Marianas and Alaska-Aleutians (0.3 events/km), and Peru (0.2 events/km).

Japan and Tonga-Kermadec also produce the most megathrust events, at ~0.9 and ~0.7 events/km, respectively. Kuril-Kamchatka is next at 0.4 events/km, and several others group together at ~0.2–0.3 events/km (Chile, Sumatra, Mexico– Middle America, and Alaska-Aleutians). Peru and Marianas produced the fewest megathrust events in this time period, at ~0.09 and 0.08 events/km, respectively.

All

Thrustn = 187b = 0.73

n = 491b = 0.87

–82° –80° –78° –76° –74° –72° –70° –68°

–18°

–16°

–14°

–12°

–10°

–8°

–6°

–4°

log(

#Eve

nts/

year

Mag

nitu

de >

Mw

) 1.5

1.0

0.5

0.0

–0.5

–1.0

–1.5

–2.0

Mw5.0 6.0 7.0 8.04.5 5.5 6.5 7.5

log(

#Eve

nts/

year

Mag

nitu

de >

Mw

)

0.5

0.0

–0.5

–1.0

–1.5

Mw5.0 6.0 7.05.5 6.5 7.5 8.0 8.5

8.5

100 km

Figure 9. Thrust faulting earthquake distri-bution and Mw-frequency distributions for the Peru subduction zone from the Global Centroid Moment Tensor (GCMT) catalog. See description of Figure 3 for details.

Page 11: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1478Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

The regional differences in seismic productivity are also manifested in large event aftershock productivity, with lower values found in Mexico–Middle America, Peru, and Chile and higher values found in Tonga, Sumatra, Japan, and Southwest Pacific island arcs (e.g., Singh and Suárez, 1988; Wetzler et al., 2016). In general, island-arc megathrusts that host great earthquakes tend to be more productive than continental arcs.

Cumulative Seismic Moment versus Depth

We compiled cumulative seismic moment from the GCMT database for all thrust-mechanism earthquakes and for all likely megathrust events with strike aligned with the trench. Seismic moment is predominantly asso-ciated with the likely megathrust events. For Chile, Japan, and Sumatra, the three subduction zones with Mw ≥ 8.8 events during the 1976 to 2016

period, cumulative moment is largely concentrated in the 15–25 km depth range (Fig. 12). The histograms are dominated by the largest earthquakes, namely, the 2004 Mw 9.2 Sumatra-Andaman, 2010 Mw 8.8 Maule, Chile, and 2011 Mw 9.1 Tohoku, Japan, earthquakes, and the moment for these events is actually distributed over a depth range rather than at the GCMT centroid depth. The other subduction zones have less cumulative seismic moment in the GCMT time period, and we found more diversity in the depth range of peak moment in each region (Fig. 13). For example, both the Alaska and Peru subduction zones had a peak moment in the 20–30 km range, but the Mexico–Middle America zone seismic moment peaked at shallower depth, between 10 and 20  km. Both the Kuril and Tonga subduction zones have recorded multiple peaks in moment depth distribution. Kuril moment peaks have been recorded at both 5–10 km and 20–30 km depths. There is another peak in the Kuril zone that is much deeper, between 60 and 65 km, although the corresponding thrust-mechanism events have strikes inconsistent with

90° 92° 94° 96° 98° 100° 102° 104° 106° 108°

–4°

–2°

10°

12°

14°

Thrustn = 691b = 0.74

Alln = 1783b = 0.88

log(

#Eve

nts/

year

Mag

nitu

de >

Mw

) 2.0

1.5

1.0

0.5

0.0

–0.5

–1.0

–1.5

–2.0

log(

#Eve

nts/

year

Mag

nitu

de >

Mw

)

1.5

1.0

0.5

0.0

–0.5

–1.0

–1.5

–2.0

Mw4.0 5.0 6.0 7.0 8.0 9.0

Mw4.0 5.0 6.0 7.0 8.0 9.0

100 km

Figure 10. Thrust faulting earthquake dis-tribution and Mw -frequency distributions for the Sumatra-Andaman subduction zone from the Global Centroid Moment Tensor (GCMT) catalog. See description of Figure 3 for details.

Page 12: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1479Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

178° 180° –178°–176°–174°–172°–170°

–34°

–32°

–30°

–28°

–26°

–24°

–22°

–20°

–18°

–16°

Thrustn = 1739b = 0.93

Alln = 3513b = 0.92

log(

#Eve

nts/

year

Mag

nitu

de >

Mw

)

1.5

1.0

0.5

0.0

–0.5

–1.0

–1.5

–2.0

log(

#Eve

nts/

year

Mag

nitu

de >

Mw

)

0.5

0.0

–0.5

–1.0

–1.5

Mw5.0 6.0 7.05.5 6.5 7.5 8.0 8.54.5

1.5

1.5

2.0 Mw5.0 6.0 7.05.5 6.5 7.5 8.0 8.54.5

2.0

2.5

100 km

Figure 11. Thrust faulting earthquake dis-tribution and Mw-frequency distributions for the Tonga-Kermadec subduction zone from the Global Centroid Moment Tensor (GCMT) catalog. See description of Fig-ure 3 for details.

TABLE 2. REGIONAL b VALUE CALCULATIONS BASED ON GLOBAL CENTROID MOMENT TENSOR (GCMT) CATALOG DATA

Region Number of all earthquakes b value (Aki-Utsu, Mc = 5.2) Number of megathrust earthquakes b value (Aki-Utsu, Mc = 5.2)

Alaska-Aleutians 1425 0.82 893 0.79Central America 1252 0.76 659 0.71Chile 1175 0.77 721 0.73Japan 1257 0.81 720 0.75Kuril/Kamchatka 1666 0.83 1203 0.80Marianas 732 1.03 181 1.07Peru 491 0.87 188 0.73Sumatra 1783 0.88 691 0.74Tonga-Kermadec 3513 0.92 1739 0.93

Note: Depth range is 0–100 km, and time range is 1 January 1976 to 7 December 2016 for all regions. Megathrust earthquakes are thrust events with focal mechanism strike within the trench strike range listed in Table 1. The b value estimates based on maximum likelihood using Aki-Utsu equation (Aki, 1965; Utsu, 1965) are included, withmagnitude of completeness (Mc) based on formal GCMT estimate of Mc = 5.2.

Page 13: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1480Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

the local trench orientation, so they are likely intraslab events. The Tonga subduction zone includes a shallow peak at 10–15 km, and a deeper one at 40–45 km; again, the deeper one is likely to be from intraslab compressional events (e.g., Meng et al., 2015b). We believe that all of the significant inter-plate seismic moment release is at depths less than 50 km in our selected regions, and less than 60 km globally.

Depth Variations in Radiation Characteristics

In recent decades, many observations have indicated depth dependence along megathrusts for a variety of seismic characteristics. The recognition of tsunami earthquakes (Kanamori, 1972) that produced much larger tsunamis than would be expected for a typical earthquake with the same surface wave magnitude (MS) prompted focused research on depth-dependent characteris-

tics. These rare shallow ruptures tend to have anomalous properties, with rela-tively low rupture velocity, low radiated energy, and large slip that generates a strong tsunami. Such events have struck along the Aleutians, Japan, the Kuril Islands, Peru, Nicaragua, El Salvador, New Zealand, Java, and Sumatra, so they are widespread, albeit infrequent events (e.g., Kanamori, 1972; Polet and Kanamori, 2000; Bilek and Lay, 2002).

In addition, the occurrence of several great (M > 8) earthquakes since 2004 has provided excellent recordings by large numbers of broadband seismo-graphs, leading to observations of large slip but very little in the way of coher-ent bursts of high-frequency radiation in the shallowest part of the relevant subduction zone, but many bursts of high-frequency radiation originating in the deeper regions of the seismogenic zone, where there is lower coseismic slip (e.g., Lay et al., 2012). In the following sections, we outline several of the depth-dependent seismic characteristics that have been noted for megathrust earthquakes.

Dep

th (k

m)

All Thrust Thrust (Trench-Strike)

Seismic Moment (dyne-cm)

Chile

Japan

Sumatra

0

20

40

60

80

1000 1 1029 2 1029 3 1029 4 1029 5 1029 6 1029

Figure 12. Cumulative seismic moment release with depth in the subduction zones that have produced the largest events Mw ≥ 8.8 since 1976. Moment, centroid depth, and focal mechanism are from the Global Centroid Moment Tensor (GCMT) catalog. Solid lines indicate those thrust mechanism events with focal mechanism strikes that fall within the regional trench strike listed in Table 1; dashed lines represent moment from all thrust mechanism earthquakes in the speci-fied region (which mostly overlap indistinguishably for these regions).

Dep

th (k

m)

Alaska

All Thrust Thrust (Trench-Strike)

Central AmericaKuril

MarianasTonga

Peru

Seismic Moment (dyne-cm)

0

20

40

60

80

1000 2 1028 4 1028 6 1028 8 1028 1 1029

Figure 13. Cumulative seismic moment release with depth in the indicated subduction zones. Moment, centroid depth, and focal mechanism are from the Global Centroid Moment Tensor (GCMT) catalog. Solid lines indicate those thrust mechanism events with focal mechanism strikes that fall within the regional trench strike listed in Table 1; dashed lines represent moment from all thrust mechanism earthquakes in the specified region.

Page 14: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1481Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

Source Duration

Tsunami earthquake ruptures have significantly longer rupture durations than is typical for similar-sized events. For example, the 1992 Mw 7.7 Nicaragua tsunami earthquake had source time function duration estimates of 100–150 s (e.g., Kanamori and Kikuchi, 1993; Ihmlé, 1996b), and the 2010 Mw 7.8 Men-tawai event had duration estimates of 80–125 s (Yue et al., 2014a; Newman et al., 2011).

Motivated by these long source duration observations for the tsunami earthquakes, Bilek and Lay (2002) examined over 500 earthquakes (Mw 5.0–7.5) within 14 circum-Pacific subduction zones and found that some events, but not all, in the shallowest 15 km of the megathrust had long moment-scaled source durations (to an equivalent Mw = 6.0 reference), similar to scaled dura-tions of seven tsunami earthquakes. Results for individual subduction zones indicate several events with long durations (10–20 s) located in the uppermost 15 km, with events deeper than 15 km having shorter (mean ~5–6 s) durations. El Hariri et al. (2013) examined relocated subduction zone earthquakes with improved depth constraints and classified the events into shallow (<26 km) and deep (>26 km) groups based on overall longer moment-scaled durations of the shallow group. El Hariri et al. (2013) also focused on regional compari-sons, finding weak to no correlation between longer duration and reported amounts of subducted sediment, presence of bathymetric features, or regions of observed afterslip, all factors that have been proposed to affect rupture complexity and duration.

More recent work suggests some depth variation in duration estimates us-ing global data sets. Denolle and Shearer (2016) examined spectral ratios for over 900 thrust earthquakes (M > 5.5) and found weak dependence of scaled source duration with increasing depth on the fault. Similarly, Ye et al. (2016c) found longer-duration earthquakes in the shallowest (<18 km) portion of the subduction megathrust using finite fault determinations for over 100 Mw > 7 earthquakes. Tsunami earthquakes, as well as a few other shallow events, were found to have scaled durations (to an equivalent Mw = 6.0 reference) of between ~8 and 19 s, i.e., significantly larger than the average of the rest of the events in their study (~5.6 ± 1.5 s).

mb/Mw

Recent observations of depth dependence in high-frequency radiation (e.g., Lay et al., 2012) suggest that other seismic characteristics, such as the short-period body wave magnitude (mb), may also vary with depth relative to the long-period Mw. Tsunami earthquakes, given their shallow slip with limited high-frequency radiation, can be identified by their relatively low mb relative to Mw measurement (e.g., Kanamori, 1972; Kanamori and Kikuchi, 1993; Lay et al., 2011a). Rushing and Lay (2012) examined the differences between mb and Mw for a large catalog of likely megathrust events with Mw ≥ 5 in several subduction zones (Chile, Japan, Sumatra, Kuril, Aleutians, Sumba, Peru). They

found positive mb-Mw perturbations for deeper megathrust events, reflecting enhanced deep short-period radiation in Japan, northern Sumatra, and central Chile. Little to no apparent depth trend was observed in the scattered data for the other subduction zones in their study.

Spatial Variations of Slip

Recent well-recorded large and great earthquakes have displayed exten-sive diversity in rupture characteristics, with differences in radiated energy depending on depth of the rupture patch, both bilateral and unilateral rup-ture expansion, and highly variable slip distribution complexity. Hundreds of events in subduction zones now have finite-fault solutions, derived from a mix of teleseismic, strong motion, geodetic, and tsunami observations. The USGS-NEIC routinely computes high-quality solutions for large shallow earthquakes (Hayes, 2017), and many contributions are collected in the SRCMOD database (http:// equake -rc .info /SRCMOD/). Some solutions are well resolved, others have severe trade-offs between kinematic constraints and model parameter-ization. Point source time functions are routinely determined by the SCARDEC method (e.g., Vallée and Douet, 2016). Efforts are under way to synthesize the accumulated data sets (e.g., Kanamori, 2014; Lay, 2015; Ye et al., 2016b, 2016c; Denolle and Shearer, 2016; Meier et al., 2017; Melgar and Hayes, 2017; Hayes, 2017), and we will not attempt to summarize the multitude of studies. Large shallow coseismic slip occurred in the 2015 Illapel (e.g., Li et al., 2016; Melgar et al., 2016), 2011 Tohoku (e.g., Lay et al., 2011b; Iinuma et al., 2012; Ozawa et al., 2012; Satake et al., 2013; Romano et al., 2014; Bletery et al., 2014; Melgar and Bock, 2015; Lay, 2017), 2010 Maule (e.g., Vigny et al., 2011; Yue et al., 2014b; Yoshimoto et al., 2016; Maksymowicz, et al., 2017), and 2004 Sumatra (e.g., Ammon et al., 2005; Rhie et al., 2007; Fujii and Satake, 2007) events, accompa-nying slip on the downdip portions of the megathrusts. In other cases, such as the 2014 Iquique, Chile (e.g., Lay et al., 2014; Hayes et al., 2014b), 2012 Nicoya, Costa Rica (e.g., Yue et al., 2013; Liu et al., 2015), 2003 Tokachi-oki, Japan (e.g., Miyazaki and Larson, 2008; Romano et al., 2010), and 2007 Pisco, Peru, rup-tures (e.g., Lay et al., 2010a; Sladen et al., 2010), slip was concentrated on the central or deeper portion of the rupture zone, with no shallow coseismic slip.

Several subduction zones have experienced aseismic slip transients both before and after the main shocks. The 2011 Tohoku and 2014 Iquique earth-quakes had large numbers of migrating repeating earthquakes during active foreshock sequences, with the repeating earthquakes providing a proxy for shallow aseismic slip that migrated toward the eventual main-shock hypo-centers (e.g., Kato et al., 2012; Ruiz et al., 2014; Hayes et al., 2014b; Kato and Kakagawa, 2014; Meng et al., 2015a). Socquet et al. (2017) describes GPS data analysis that supports the presence of a slow slip event in the 2014 Iquique re-gion during the 8 mo prior to the main shock. Aseismic slip triggered by great earthquakes is also common (e.g., Chlieh et al., 2007; Hsu et al., 2006; Ozawa et  al., 2011; Hayes et  al., 2014a; Uchida et  al., 2015). Aseismic, or slow-slip, transients are not limited to regions of recent great earthquakes, as these have

Page 15: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1482Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

been observed in both shallow and deep portions of the seismogenic zone in many subduction zones around the globe (e.g., Dragert et al., 2001; Obara et al., 2004; Brudzinski et al., 2007; Schwartz and Rokosky, 2007; Correa-Mora et al., 2009; Jiang et al., 2012; Saffer and Wallace, 2015; Wallace et al., 2016).

Many of the great earthquakes in the last 39 yr have occurred in previ-ously recognized seismic gaps (McCann et al., 1979; Nishenko, 1991), although some only partially filled the gap, leaving residual strain accumulation to be released in future events (Fig. 14). Targeted instrumentation of specific iden-tified seismic gaps prior to great earthquakes in 2010, 2014, and 2015 along Chile, in 2005 and 2007 along Sumatra, in 2016 along Ecuador, and in other locations, has resulted in unprecedented seismic and geodetic data sets for recent events. However, even given the general constraint of the plate-bound-

ary slip budget, the diversity of recent large earthquakes on megathrusts has presented surprises in terms of event location, size, and hazard for multiple plate boundaries. The 2004 Sumatra earthquake ruptured over a much greater length (>1300 km) of the plate boundary than had been considered likely (e.g., Shearer and Bürgmann, 2010), and the 2011 Tohoku, Japan, earthquake rup-tured with unexpectedly huge slip, extending to the trench, along with rerup-turing regions of recent, smaller earthquakes (e.g., Lay and Kanamori, 2011; Fujiwara et  al., 2011; Iinuma et  al., 2012; Lay, 2017). Both events generated devastating tsunamis and reaffirmed the potential for events larger than have been documented in recorded history to strike in various regions.

The 2015 Illapel earthquake ruptured the Chilean megathrust close to previ-ous earthquakes in 1880 and 1943, and so the 2015 event has been suggested

Figure 14. Seismic gaps, as defined by McCann et  al. (1979), and recent great earthquakes (1979–2016). Colored lines indicate timing of great earthquake occur-rence prior to gap definitions in 1979, with red shading indicating longest time since great earthquake. Focal mechanisms of great earthquakes between 1979 and 2016 are from the Global Centroid Moment Ten-sor (GCMT) catalog.

Page 16: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1483Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

to be a “characteristic earthquake” for this segment of the margin (e.g., Tilman et al., 2016; Klein et al., 2017). In contrast, the 2014 Iquique earthquake spanned only about 20% of the northern Chile seismic gap, with most of the 1877 M ~ 8.8 earthquake rupture zone remaining unbroken (e.g., Hayes et al., 2014b; Meng et al., 2015a; Cesca et al., 2016). Earthquakes along northern Japan were common in the last century, mainly in the M 7–8 magnitude range, but the 2011 Mw 9 Tohoku earthquake ruptured a region where the last giant event occurred in A.D. 869, along with multiple rupture zones of previous individual ruptures (e.g., Koper et al., 2011; Lay, 2015). The 2010 Mw 8.8 Maule earthquake ruptured a seismic gap in a region that last ruptured in 1835 in an M ~8.5 earthquake (e.g., Moreno et al., 2010), but the main slip extended beyond that rupture, overlapping the rupture zone of a large event in 1928, but not the adjacent 1985 rupture zone. Along the Sumatra megathrust, the great earthquakes in 2005 and 2007 reruptured regions that had previously broken in 1797, 1833, 1861, and 1907 (e.g., Lay et al., 2005; Konca et al., 2008; Chlieh et al., 2008). The 2007 Peru earthquake occurred in a region last possibly ruptured in 1687 or 1746 and only partially filled a seismic gap between the 1996 and 1974 events (Pritchard and Fielding, 2008). The 2007 Solomon Islands earthquake ruptured in a very complex triple-junction plate-boundary zone, with no prior record of large earthquakes (Taylor et al., 2008; Furlong et al., 2009; Chen et al., 2017). Along the central Kuril Islands, the 2006 Mw 8.3 earthquake ruptured a portion of the megathrust between the 1963 Mw 8.5 and 1952 Mw 9 earthquakes identified as a seismic gap that may have partially failed in 1918 (e.g., Lay et al., 2009).

Spectrum of Slip Velocities

Observations of slip processes for subduction zone megathrusts have ex-panded greatly beyond early measurements of earthquake rupture front ve-locities, which were on the order of 75%–95% of the shear wave velocity in the slip zone expansion of the earthquake (e.g., Kanamori and Brodsky, 2004). Tsu-nami earthquakes appear to have very low rupture velocities, such as the 1–1.5 km/s rupture velocity of the 1992 Nicaragua tsunami earthquake (e.g., Kikuchi and Kanamori, 1995; Ihmlé, 1996a, 1996b) and the 1.25–1.5 km/s rupture veloc-ity for the 2010 Mentawai event (e.g., Lay et al., 2011a; Newman et al., 2011). Kanamori et al. (2010) and Bilek et al. (2011, 2016) suggested that other events in the shallow region of tsunami earthquakes may also exhibit low rupture velocities. Variable rupture speeds may exist within an individual earthquake, such as the 2010 Maule (Kiser and Ishii, 2011) event, and rupture velocities may lie along a wide continuum between “typical” and “slow” speeds, such as the ~1.5–2 km/s rupture velocity estimated for the 2013 Santa Cruz Islands earth-quake (Lay et al., 2013; Hayes et al., 2014a). Supershear earthquakes have not been observed in the shallow portion of subduction zones, possibly because the events tend to have mixed mode ruptures.

At rupture velocities much slower than for tsunami earthquakes, a fam-ily of slip processes, including low-frequency earthquakes (LFEs) and very low-frequency earthquakes (VLFEs), involves slip events with velocities or-ders of magnitude lower than typical earthquake rupture velocities (Ide et al.,

2007a). LFEs are deficient in high-frequency energy, with dominant signals around 1–8 Hz, durations of ~0.3 s, and magnitudes of ~1011 Nm, with locations and mechanisms reflecting slip on the megathrust (e.g., Shelly et al., 2007; Ide et al., 2007b). In many cases, these LFEs occur as swarms that comprise much of the seismic nonvolcanic tremor observed in subduction zone settings (e.g., Ide et al., 2007b). LFEs were initially detected and located on the deeper portion of seismogenic zones (e.g., Katsumata and Kamaya, 2003; Shelly et al., 2006; Brown et al., 2009; Bostock et al., 2012), but they have now also been observed in shallow portions of accretionary prisms (e.g., Obana and Kodaira, 2009; Yama shita et al., 2015; Nakamura, 2017).

VLFEs represent even slower slip, with durations of ~20 s, magnitudes up to ~1014 Nm, and frequencies of ~0.02–0.05 Hz, i.e., even more deficient in high-frequency energy than LFEs (e.g., Ide et al., 2007a). Similar to LFEs, these events appear to be located throughout the seismogenic zone (Fig. 1), from the deep transition between the seismogenic and aseismic zones (e.g., Ito et al., 2007, 2009) to the updip end of the seismogenic zone, in some cases, very close to the trench (e.g., Ito and Obara, 2006; Asano et al., 2008; Matsuzawa et al., 2015; Araki et al., 2017), and within the accretionary prism (e.g., Obara and Ito, 2005; Ito and Obara, 2006).

At the slowest end of the slip spectrum, there are slow slip events (SSEs), which are observed geodetically along many subduction zones. These slip events can be quite large, reaching an equivalent seismic moment of 1018–1020 Nm, with durations over 105–107 s (days to months; e.g., Ide et  al., 2007a). SSEs are often, but not always, accompanied by nonvolcanic tremors and are grouped into the episodic tremor and slip (ETS) category (e.g., Dragert et al., 2001; Kostoglodov et  al., 2003; Obara et  al., 2004; Hirose and Obara, 2006; Schwartz and Rokosky, 2007; Beroza and Ide, 2011). These were also initially observed in the vicinity of the deepest portions of the seismogenic zone (Fig. 15), but more recent observations indicate that shallower megathrust SSEs occur as well (Saffer and Wallace, 2015, and references therein), including very close to the trench (Araki et al., 2017).

Radiated Energy, Apparent Stress, and Stress Drop

Choy and Boatwright (1995) determined radiated energy from global shal-low earthquakes (M > 5.8) and used it to compute apparent stress (given by the ratio of radiated energy to seismic moment, ER /M0, multiplied by the average rigidity). They found that subduction zone thrust events have the lowest val-ues (~0.29 MPa) of apparent stress for any tectonic category, although with re-gional variations between 0.15 MPa (Kermadec) and 0.42 MPa (Peru-Ecuador).

Newman and Okal (1998) determined radiated energy and energy-to- moment ratios for over 50 teleseismic earthquakes, highlighting the 1992 Nica ragua, 1994 Java, and 1996 Peru tsunami earthquakes as having distinctly low values for moment-scaled radiated energy relative to other earthquakes in their data set. Okal and Newman (2001) examined other earthquakes within these three tsunami earthquake-producing regions, finding additional “slow” earthquakes present only near the 1960 Peru tsunami earthquake zone.

Page 17: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1484Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

Convers and Newman (2011) expanded the catalog of Newman and Okal (1998) by determining estimates of radiated seismic energy for 342 earthquakes from 1997 to 2010, finding shallow subduction zone thrust earthquakes generally deficient in radiated energy. They found additional energy-deficient events in the rupture zones of the 1992 Nicaragua and 2006 Java tsunami earthquakes. Venkataraman and Kanamori (2004) and Ye et al. (2016c) also demonstrated that tsunami earthquakes have low moment-scaled radiated energy compared to other subduction zone earthquakes.

For events within a given region, recent great earthquakes that have been well recorded by large seismic networks have a depth dependence in the frequency of radiated seismic energy (e.g., Lay et al., 2012). The deeper parts of the megathrust produce significantly more high-frequency energy than the shallower megathrust for most of the largest earthquakes in recent decades, such as the 2010 Maule Chile Mw 8.8 (e.g., Kiser and Ishii, 2011; Koper et al., 2012), 2011 Tohoku Mw 9.1 (e.g., Koper et al., 2011; Ide et al., 2011; Ishii,

2011), and the 2015 Mw 8.3 Illapel, Chile, event (e.g., Melgar et al., 2016; Yin et al., 2016). Ye et al. (2016c) documented modest depth dependence in high- frequency radiation for the 100+ large events, with lower spectral decay rates for deeper events on subduction zone megathrusts.

Allmann and Shearer (2009) determined stress drop for over 1700 global earthquakes, finding little depth variation in the full global catalog, but region-ally significant depth-dependent stress drops. They also found along-strike variations, in particular, very low values along portions of the Central America subduction zone and near the hypocenter of the 2004 Sumatra earthquake. Similar to the depth dependence of source duration discussed in the “Source Duration” subsection herein, Denolle and Shearer (2016) found very weak depth dependence for stress drop and scaled energy using their global cata-log. They observed regional variations, with lower stress drop in general in western Pacific subduction zones as compared to eastern Pacific zones, and very low stress drops for earthquakes along the Alaska-Aleutian subduction

10203040506070km

0

200

400

km

0

200

400

km

0

200

400

km

0

200

400

km

0

200

400

km

0 200 400 600km

−156°

−154°

−152°

–150°

−148°

−146°

−144°

−142°

58°

58°60°

0 200 400 600 800 1,000 1,200km

–128°

–126°

–124°

40°

42°

44°

46°

48°

50°

Alaska

CascadiaNankai

aciRatsoCocixeM

Tremor SSE Contours of plate interfaces

Megathrustearthquake 5 cm year–1

0 200 400 600 800 1,000km

130°

132°

134°

136°

138°

140°

32°

32° 34°

34°

Boso

0 200 400 600km

−102°

−100°

−98°

18°

20°

Guerrero Oaxaca

0 200 400 600km

−88°

−88°

−86°

−84°8°10°

12°

1964

Shikoku KiiTokai

1923Kanto

1946 Nankai 1944 Tonankai

Figure 15. Examples of tremor and slow-slip events (SSE) located at the downdip portion of the seismogenic zone, modified from Beroza and Ide (2011).

Page 18: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1485Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

zone. Ye et al. (2016c) found little depth dependence of stress drop for major and great megathrust events deeper than 15 km, but there is an order of mag-nitude scatter in stress drop values with little regional coherence. Our compi-lations of these various global studies also suggest weak depth dependence in both ER /M0 and stress drop, with large scatter (Figs. 16 and 17).

On regional scales, spatial variations in stress drop for small earthquakes are observed and may be useful for probing the stress state of a megathrust. Uchide et al. (2014) described stress drops computed for over 1500 small earth-quakes within the Tohoku region of Japan prior to the 2011 Mw 9.1 earthquake. They found increasing stress drop at 30–60 km depths, and strong along-strike

variations that appeared to link areas of high-stress-drop events with zones of little coseismic slip in 2011, south of the region of maximum main-shock slip. Bilek et al. (2016) used spectral ratio techniques to compute corner fre-quencies, seismic moment, and stress drop for a series of earthquakes that occurred within and adjacent to the 1992 Nicaragua tsunami earthquake rup-ture zone. Earthquakes from 2005 to 2006 that occurred within the 1992 rupture zone exhibited significantly lower corner frequencies and stress drop than those events that occurred adjacent to the 1992 rupture zone, suggesting that regional variations in the megathrust zone that affect rupture processes may persist for many years afterward, if not permanently.

10–8

10–7

10–6

10–5

0 10 20 30 40 50 60 70 80

Denolle and Shearer (2016)Ye et al. (2016, group 2 events)Ye et al. (2016, group 1 events)

Depth (km)

Er/

Mo

10–4

10–3

10–2

Figure 16. Radiated energy (Er) over seismic moment (MO) estimates for global subduction zone earthquake populations from Denolle and Shearer (2016) and Ye et  al. (2016b). The Ye et  al. (2016b) study included only megathrust ruptures; group 1 events are those with independent rupture model constraints; rupture characteristics for group 2 events were determined using a range of rupture velocities in kinematic inversions of teleseismic body waves.

0.001

0.01

0.1

1

10

100

1000

104

0 10 20 30 40 50 60 70 80S

tres

s D

rop

(MP

a)

Depth (km)

Ye et al. (2016, group 1 events)Ye et al. (2016, group 2 events, Vr 3.0 km/s)Denolle and Shearer (2016)El Hariri et al. (2013)

Figure 17. Stress drop estimates for global compilations of shallow subduction zone earth-quakes from Ye et al. (2016b), Denolle and Shearer (2016), and El Hariri et al. (2013). Stress drop estimates from Ye et al. (2016b) include the group 2 event rupture models with a Vr of 3 km/s.

Page 19: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1486Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

Aftershock Distribution

Compilations of aftershock locations relative to areas of high slip during large earthquakes of all types suggest that aftershocks tend to locate at bound-aries between areas of high slip and low slip and/or at edges of large-slip re-gions (e.g., Das and Henry, 2003; Wetzler et al., 2018). This observation has been supported by aftershock catalogs for many of the M > 8 earthquakes, such as the recent Sumatra (e.g., Hsu et al., 2006), Chile (Hayes et al., 2014b; Yue et al., 2014b; Li et al., 2016), and Tohoku (e.g., Asano et al., 2011) earth-quakes. The 2013 Santa Cruz Islands earthquake appears to have had few after-shocks on the megathrust fault, and this event may have triggered an adjacent aseismic slip episode following the main shock (Hayes et al., 2014a). The 2014 Iquique earthquake produced many aftershocks, but the aftershocks had low b values, despite the occurrence of a large aftershock (Hayes et al., 2014b). There is a weak tendency for high-stress-drop megathrust ruptures with a given mo-ment to have lower aftershock production, which is attributed to the smaller main-shock rupture dimensions (Wetzler et al., 2016). The source dimensions of the main slip zone for the 2014 Iquique event were indeed unusually small for a great earthquake.

Aftershocks occur off of the main megathrust rupture plane as well. Nor-mal faulting aftershocks have been observed after many large megathrust rup-tures (Fig. 18), particularly those that rupture very shallowly near a trench (e.g., Kanamori, 1971; Christensen and Ruff, 1988; Lin and Stein, 2004; Ammon et al., 2008; Lay et al., 2009; Asano et al., 2011; El Hariri and Bilek, 2011; Bilek et al., 2011; Ide et al., 2011; Kato et al., 2011; Lange et al., 2012; Rietbrock et al., 2012; Yue et al., 2014b; Wetzler et al., 2017; Sladen and Trevisan, 2018). Some of these intraplate extensional events are within the outer rise, seaward of the trench, or on plate-bending related faults below the megathrusts, while others are found within the upper plate adjacent to the zone of high slip (e.g., Farías et al., 2011; Hicks and Rietbrock, 2015). Events with rupture deeper on the mega thrust activate relatively few intraplate aftershocks (Wetzler et al., 2017; Sladen and Trevisan, 2018), but in some cases, they have large afterslip on the updip re-gion of the megathrust (Fig. 18).

FACTORS CONTRIBUTING TO MEGATHRUST SPATIAL HETEROGENEITY AND SLIP VARIABILITY

As demonstrated in the previous section, earthquake observations now span a wide range of slip speed and complexity. Here, we outline a range of possible factors leading to these diverse slip observations.

Key questions about where slip can occur in subduction zones often start with an examination of seismic coupling. This can be defined and estimated in a variety of ways (e.g., Wang and Dixon, 2004), but it fundamentally involves a short time window estimate of how much of the plate motion is released in fast processes that generate seismic energy (e.g., Scholz and Campos, 2012). Var-ious studies have estimated seismic coupling for large segments of subduc-

tion zones in order to compare it with other parameters, such as slab age and dip, to understand the large-scale controls on coupling. With the availability of longer earthquake catalogs, more detailed seismic observations of individual earthquakes, and increasingly widespread geodetic observations, other stud-ies have focused on smaller spatial scales to examine coupling and the rela-tionship to seismic activity. We summarize both scales of observations here.

Large-Scale Subduction Zone Parameters

Uyeda and Kanamori (1979) described end-member subduction models with distinct back-arc processes. They noted that zones with active back-arc spreading, such as the Marianas, have fewer large and great earthquakes than zones without back-arc spreading, such as the great earthquake-producing Chilean margin. As an explanation for these end-member behaviors, they sug-gested each zone has a different level of plate coupling caused by either time evolution of the subduction geometry and progressive decoupling or differ-ences in the motion of the overriding plate.

In their broader review of plate coupling in many subduction zones around the globe, Scholz and Campos (1995) suggested that the strongest coupling occurs where large megathrust normal forces exist, controlled by plate age, slab length and dip angle, and velocity of plate motion. Subduction zones with active back-arcs, and old, steeply dipping subducting lithosphere such as the Marianas, southern Tonga, and Izu-Bonin have the lowest normal force acting on the interface in this model, and so they have the weakest coupling and smaller maximum-magnitude earthquakes. In contrast, the Chile, Peru, Sumatra, Alaska, and Cascadia subduction zones have higher normal forces and plate coupling. Scholz and Campos (2012) revisited this model with up-dated seismic catalogs and coupling estimates finding a similar correlation, with the larger earthquakes occurring in regions of higher normal forces and higher coupling. They also took advantage of the large number of geodetically based coupling estimates to resolve significant along-strike coupling varia-tions within individual subduction zones.

Because of the importance of slab dip angle on normal forces and coupling, various studies have examined dip angles in the context of other subduction zone parameters. For example, a comparison of slab dip in over 150 geomet-rically simple subduction zone transects suggests that the absolute motion of the overriding plate has the strongest correlation with slab dip, much larger than factors such as the age and thermal structure of the slab, the slab pull force, or convergence rate (Lallemand et al., 2005). Back-arc spreading is only found in areas where the deeper portion of the slab dips at greater than 50°, and back-arc shortening is prevalent in areas with slab dips of less than 31° (Lallemand et  al., 2005). Curvature of the megathrust is strongly correlated with average dip angle of the megathrust, but it is also influenced by dip of the deeper slab, which may cause shear strength to vary across the plate bound-ary. More shallowly dipping, flatter boundaries are able to support larger earthquakes due to more homogeneous strength (e.g., Bletery et  al., 2016).

Page 20: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1487Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

A

B

C

Figure 18. Examples of range of outer- rise seismicity characteristics follow-ing large and great megathrust earth-quake ruptures, reprinted from Sladen and Trevisan (2018), with permission from Elsevier. In several large to great earthquakes with slip distributions (green-blue shading) indicating slip to the trench, significant numbers of normal faulting aftershocks occurred (red circles). In areas with little to no rupture to the trench, normal-faulting aftershocks were less prevalent. Global CMT—Global Centroid Moment Tensor catalog. ISC, International Seismologi-cal Centre; WR, Wharton Ridge. Events such as HaidaGwaii and Tocopilla are discussed in Sladen and Trevisan (2018).

Page 21: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1488Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

The nature of the upper plate, being either continental or oceanic, also appears to be important, as are factors such as the presence of back-arc spreading and trench roll-back (Scholz and Campos, 2012).

Role of Megathrust Fluids

Fluids play a critical role in many subduction zone processes, from fault slip to volcanic activity (e.g., Peacock, 1990). We focus here on the aspects of fluids that influence slip processes on the megathrust, from great earthquakes to the slow-slip and tremor episodes observed in many regions. Fluids enter the shallow seismogenic system either within the pore space of the subducting igneous crust and sediment, or bound in hydrous minerals. Normal faulting associated with bending of the subducting plate is likely to produce enhanced hydration of the crust and upper mantle seaward of the subduction zone (e.g., Ranero et al., 2003; Naliboff et al., 2013; Korenaga, 2017; Grevemeyer et al., 2018). These fluids are released either through compaction in the shallowest 5–7 km or in dehydration reactions at higher pressures and temperatures (e.g., Saffer and Tobin, 2011). The fluids move through the system by a variety of pathways, including faults, fractures, and permeable strata within the mega-thrust and overriding plate (e.g., Carson and Screaton, 1998), or they remain trapped by low-permeability sediments, affecting the seismogenic zone by increasing fluid pressures (e.g., Saffer and Bekins, 2002). Excess fluid pres-sures within the megathrust zone modify the effective normal stress, which is linked to overall fault strength and generation of both typical earthquakes and slower- slip processes (e.g., Saffer and Tobin, 2011).

Role of Fluids in Seismic Slip

Because high fluid pressure acts to reduce effective normal stress, fluids play an important role in the earthquake process. Various studies have pro-vided evidence for significant fluid content and high fluid pressures in the megathrust zone, often based on bright spots in seismic reflection profiles and high Vp/Vs seismic velocity ratios, which serve as a proxy for high fluid pres-sures. We note only a few of these studies here, along with the seismic implica-tions of these observations. Tobin and Saffer (2009) estimated high pore-fluid pressures and low effective stress along the shallow Nankai subduction zone using detailed seismic reflection data. Similarly, Park et al. (2014) attributed along-strike variability in seismic reflection characteristics along the Nankai subduction zone to differences in fluid amounts along the megathrust. Bangs et al. (2015) used three-dimensional (3-D) seismic imaging to identify fluid-rich segments along the plate interface offshore central Costa Rica. Audet et al. (2009) used seismic receiver functions and Vp/Vs to suggest that the mega-thrust zone along the northern Cascadia margin has low permeability, acting to seal high fluid pressures of the oceanic crust, and leading to significant over-pressures in this zone. Audet and Schwartz (2013) found along-strike variations

in upper-plate Vp/Vs and inferred fluid pressure variations along a small seg-ment of Costa Rica, with lower Vp/Vs in the northwestern portion of the Nicoya Peninsula relative to the higher Vp/Vs and higher fluid pressures along the southeastern segment of the peninsula. They suggested that these upper-plate variations influence the seismogenic zone behavior.

These variations in fluid content and pressure are also linked to variations in slip processes. Along the Nankai subduction zone, the high fluid pressures in the shallow zone may inhibit near-trench slip (Tobin and Saffer, 2009), and the along-strike variability at depth may allow for nucleation of tsunami earth-quakes in fluid-poor regions, with rupture into fluid-rich zones later in the event (Park et al., 2014). Along central Costa Rica, the seismicity appears to preferentially occur in the fluid-poor zones (Bangs et al., 2015), and along the Nicoya Peninsula, Costa Rica, the areas of low fluid pressures host small-mag-nitude earthquakes with higher apparent stresses than are found in the high-fluid- pressure areas (Audet and Schwartz, 2013; Stankova-Pursley et al., 2011).

Other studies have compared seismic observations of Vp/Vs and fluid pres-sure estimates to geodetic coupling and rupture details of specific earthquakes. For example, Moreno et al. (2014) examined a portion of the 2010 Maule Chile rupture zone and found a positive correlation between areas of high Vp/Vs and low geodetically determined coupling. Huang and Zhao (2013) and Yamamoto et al. (2014) found significant Vp/Vs heterogeneity in the shallow megathrust zone of the 2011 Tohoku rupture, suggesting a high Vp/Vs ratio in the very near-trench area of high coseismic slip, as well as downdip of the hypocenter of the event, with low Vp/Vs just updip of the hypocenter.

Slow-Slip and Tremor Processes

Many observations of slow slip and tremors in subduction zones corre-spond to areas that appear to have high fluid pressures. Early efforts to iden-tify and locate nonvolcanic tremors in the Nankai subduction zone resulted in locations at depths where pressures and temperatures are such that fluids can be generated through dehydration (Obara, 2002). Rogers and Dragert (2003) emphasized the potential importance of fluids in their initial reporting of Cas-cadia episodic tremor and slip. Tremor signals along the Nicoya Peninsula, Costa Rica margin, correlate with observations of seafloor fluid-flow measure-ments at the surface of the forearc, which is explained by slow-slip motions in the shallow portion of the megathrust causing the observed transient flow in the upper-plate sediments (Brown et al., 2005).

LFEs have also been linked to high fluid pressures. Shelly et al. (2006) lo-cated LFEs within tremors along the Nankai subduction zone in an area of high Vp/Vs, indicative of high pore-fluid pressures. Recent locations of LFEs in Nan-kai correlate variations in LFE location with Vp and Vs anomalies. Nakajima and Hasegawa (2016) found gaps in LFE occurrence in areas where they ob-served small Vp and Vs anomalies, indicative of a well-drained and low-pore-pressure megathrust; the LFEs are instead concentrated in areas of higher Vp and Vs anomalies, serving as proxies for higher-pore-fluid pressure.

Page 22: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1489Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

Similar correlations exist for slow-slip events and high fluid pressures. Seis-mic images suggestive of a high-fluid-pressure zone in the area of slow earth-quakes along the Nankai subduction zone prompted Kodaira et al. (2004) to link the slow-slip event generation with fluids. Song et al. (2009) documented a very low-seismic-velocity layer in the same region as several slow-slip events along the Mexico subduction zone; they postulated that the low velocities are the result of a region with high pore-fluid pressure. Seismic reflection data suggest high fluid pressures in slow-slip zones of the Hikurangi margin (Bell et al., 2010). Audet and Schwartz (2013) identified similar correlations among high Vp/Vs, higher fluid pressures, and slow slip in the southeastern Nicoya segment of the Costa Rica subduction zone. The high fluid pressures, and the related decrease in the effective normal stress are commonly invoked as nec-essary for these slow-slip processes in the both the shallow (e.g., Saffer and Wallace, 2015) and deep portions of the seismogenic zone (e.g., Audet and Kim, 2016).

Observations of finer-scale spatial and temporal distribution of fluid pres-sure are now being made and related to slow slip and tremor processes. Kato et al. (2010) found an asymmetric pattern of highest Vp/Vs in the area of slow slip along the Tokai segment of Japan, with lower Vp/Vs and fluid pressures in the deeper segment of the fault where the LFEs and tremor occur. Frank et al. (2015) suggested variable fluid pressures over fairly short time and spa-tial scales based on acceleration and deceleration of LFE activity in Mexico adjacent to motion during the 2006 slow-slip event.

Role of Sediments on Megathrusts

Another important input into the subduction zone is sediment, as the quan-tity and mineralogical content can affect shallow slip processes. Comparisons between locations of great earthquakes and sediment thickness highlight the first-order observation that great earthquakes occur in regions with thick sediment inputs (Ruff, 1989). This observation led to the idea that thick sub-ducted sediment packages act to smooth the plate-boundary zone, resulting in uniform seismic coupling that allows earthquakes to rupture unimpeded for large distances on the megathrust (e.g., Ruff, 1989). This model has per-sisted throughout the last three decades of seismicity and improved sediment thickness estimates (e.g., Heuret et al., 2012), with Scholl et al. (2015) finding the majority of great earthquakes occur in subduction zone segments with sediment thickness greater than 1 km (Fig. 19). Seno (2017) observed that a thickness of subducted sediment greater than 1.3 km is found for regions with Mw ≥ 9.0 earthquakes, and such events do not occur if the thickness is less than 1.2 km.

As with all of these factors, spatial variations in sediment thickness may be relevant for understanding the details of slip patterns. Various diagenetic and metamorphic processes that affect the friction and effective stress in the shallow megathrust materials play important roles in defining the updip edge of seismicity (Moore and Saffer, 2001). Thick and strong sediments in the 2004

Sumatra earthquake rupture zone may have allowed very shallow rupture for that event, in contrast to the deeper slip of the 2005 Nias earthquake (Geersen et al., 2013). Large variations of subducted sediment along-strike in northern Hikurangi may affect the locations of shallow slow-slip events there (Bell et al., 2010; Eberhart-Phillips and Bannister, 2015). Based on determination of seismic velocities in the location of 2011 Tohoku rupture, Zhao et al. (2011) found that high coseismic slip and foreshock locations are concentrated in the high-seis-mic-velocity areas; the nearby low-seismic-velocity patches, inferred to be sed-iments and high fluid content, host aseismic creep and slow thrust events. Be-yond the sediment thickness and composition, fluids expelled from subducted sediments also appear to be an important component in the slow-slip process (e.g., Peacock, 2009; Peng and Gomberg, 2010; Audet and Kim, 2016).

Sediment type is also important for controlling slip behavior. Many stud-ies have focused on the frictional stability of various clay-rich and carbonate sedi ments observed entering subduction zones to identify the conditions under which velocity weakening behavior is possible (e.g., den Hartog et al., 2012; den Hartog and Spiers, 2013; Ikari et al., 2013; Kurzawski et al., 2016). Geersen et al. (2013) inferred that the clay-rich sediment composition in the 2004 Sumatra earthquake rupture zone was as important as the thickness in controlling the shallow rupture. Ikari et al. (2015a, 2015b) demonstrated that samples of the smectite-rich fault material obtained from the shallow drilling of the 2011 Tohoku earthquake zone can deform in the laboratory at a variety of rates, from fast seismic slip to slow SSE speeds, to produce the range of slip behaviors observed in that zone.

Role of Rough Downgoing Plate Topography

Yet another important input into the subduction zone is the overall rough-ness of the subducting plate. This can be smoothed with a thick sediment blan-ket, or it can be rough, marked with seamounts, ridges, and other topographic features. As mentioned in the previous section, thick sediment subduction and smooth plates have been linked with the occurrence of great earthquakes. The way in which the rough plate affects earthquake rupture appears to be more variable. From large seamounts entering the Japan Trench (e.g., Nishi-zawa et al., 2009), to smaller seamounts leaving scars in the forearc in Costa Rica (e.g., Ranero and von Huene, 2000), and large seamounts and ridges entering the subduction zone along Peru, Chile, and elsewhere (e.g., Spence et al., 1999; Robinson et al., 2006; Sparkes et al., 2010; Marcaillou et al., 2016), these features are significant enough to likely affect megathrust earthquakes, although exactly how is an active debate. Kelleher and McCann (1976) cor-related locations of great earthquakes with the relatively smooth bathymetry portions of subduction zones, finding only moderate-magnitude earthquakes occurring in areas with significant subducting bathymetry such as aseismic ridges. They suggested that subduction of these large topographic features leads to increased buoyancy at those locations, which modifies coupling in those zones, reducing the ability to produce great earthquakes. On the other

Page 23: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1490Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

end of the spectrum, Cloos (1992) suggested that many subduction zone thrust earthquakes are the result of subducted seamounts acting to strongly couple with the overriding plate, producing asperities and earthquake slip, although the ability to produce large earthquakes would reflect the amount of sediment acting as seamount cover to seismogenic depths (Cloos and Shreve, 1996).

In order to understand how subducted topography might affect earth-quakes, we need to understand the fate of the topography within the subduc-

tion zone itself. One possibility is that both normal stress and seismic coupling increase because of the extra mass and buoyancy associated with a subducted seamount (Scholz and Small, 1997). Using analogue sandbox models of sea-mount subduction, Dominguez et  al. (2000) demonstrated significant defor-mation in the upper layers of sand as the topographic feature passed at depth. Following in the spirit of the sandbox models, Wang and Bilek (2011) presented a conceptual model of strong geometric features such as seamounts signifi-

N

Thick (> 1.0 km) Trench SedimentThin (< 1.0 km)Trench Sediment

5000 km

Figure 19. Locations of magnitude ≥ 7.5 earthquakes (stars) between 1899 and 2013 relative to sediment thickness at the trench, modified from Scholl et al. (2015).

Page 24: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1491Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

cantly deforming the surrounding upper plate with a complex network of fractures, supported by geologic evidence of highly fractured exhumed upper- plate zones, seismic imaging, and recent numerical models (Ding and Lin, 2016). These models suggest different seismic behavior in areas of subducting topography, where these features could act as either regions of high slip, or regions that arrest slip during a megathrust earthquake.

The models are variously supported by a range of documented cases of these features acting either as barriers to rupture or as regions of high slip. The ideas developed by Kelleher and McCann (1976) of great earthquakes avoiding areas of seamount subduction persist with consideration of up-dated earthquake catalogs (Wang and Bilek, 2014). Instead of producing large earthquakes, several of these regions exhibit high levels of creep (low seismic coupling), which are a proposed consequence of seamount deformation of the upper-plate model (Wang and Bilek, 2014). Yang et al. (2013) used numeri-cal simulations to show that subducted seamounts can act as rupture barriers over a wide range of seamount sizes and normal stress conditions. In regions adjacent to subducted topography, several great earthquakes have been ob-served, leading to many studies suggesting that subducted topography acts as a barrier to further rupture. For example, Kodaira et al. (2000) imaged a subducted seamount at depth in the Nankai subduction zone that they suggest acted as a barrier in the 1946 Mw 8.3 Nankai earthquake rupture. Other ex-amples of significant subducted topography acting as rupture barriers include along the South American margin (e.g., Perfettini et al., 2010; Sparkes et al., 2010; Geersen et al., 2015), Sumatra margin (e.g., Chlieh et al., 2008; Henstock et al., 2016), and Japan Trench (e.g., Simons et al., 2011; Duan, 2012). In some cases, the subducting topography may initially act as a barrier and then fail later in the rupture, producing significant slip, as proposed for the 2001 Mw 8.4 Peru earthquake (Robinson et al., 2006). Given the position of the subducted topography with depth, these features might also affect the downdip earth-quake rupture behavior as well, as suggested by Marcaillou et al. (2016) as a means for impeding updip rupture of the 1942 Ecuador M 7.8 earthquake.

Other small- to moderate-magnitude earthquakes have produced slip in subducting areas with rough topography. Using high-quality seismic surveys and monitoring efforts, Mochizuki et al. (2008) located several repeating M ~ 7 earthquakes along a portion of the Japan Trench megathrust landward of a subducted seamount, not at the seamount position at depth. Several studies have suggested that subducted seamounts acted as asperities in past Costa Rica M ~ 7 earthquake ruptures (e.g., Protti et al., 1995; Husen et al., 2002; Bilek et al., 2003), M ~ 7 earthquakes along the Hikurangi margin (e.g., Bell et al., 2014), as well as the 2010 Maule, Chile, earthquake (Hicks et al., 2012).

Subducting topography has also been linked with other seismic processes, such as small-magnitude earthquake swarms, nonvolcanic tremor, and slow-slip events. Earthquake swarms along South American and Japanese mega-thrusts appear to correlate with locations of subducted topography (Holtkamp and Brudzinski, 2011). Tréhu et al. (2012) found evidence for small-magnitude earthquake clusters in areas of subducted seamounts along the Cascadia sub-duction zone. Linear streaks of nonvolcanic tremors in the Nankai subduction

zone correlate with the path of previously subducted seamounts (Ide, 2010). Bell et al. (2010) used seismic reflection images of the Hikurangi margin to suggest that subducted seamounts allow for lower effective stress and promote the oc-currence of slow slip, while Kodaira et al. (2004) suggested a similar connection between subducted ridges and slow slip in the Nankai subduction zone.

Role of Upper-Plate Structure

The upper plate may also impact megathrust slip behavior. Correlations be-tween forearc basins and the high-slip zones of many megathrust earthquakes led Wells et al. (2003) to suggest that subsidence associated with these basins might be linked with areas of subduction erosion (requiring high coupling), affecting the megathrust stress state. Similarly, correlations between large negative trench-parallel gravity anomalies and the locations of great earth-quakes may also connect forearc structure with megathrust processes (Song and Simons, 2003). Gravity anomalies are also associated with displacement in great events like the 2011 Tohoku earthquake, where positive gravity anoma-lies, representing differences in upper-plate geology, may have influenced the frictional conditions on the plate interface (Bassett et al., 2016). Geology and strength of the margin prism may also impact the amount of slip and seafloor deformation, based on numerical models showing high slip and increased tsunami generation beneath compliant prisms (Lotto et al., 2017). Upper-plate faulting may also provide limits on megathrust earthquake ruptures, as sug-gested for Ecuador (Collot et al., 2004) and Chile (Moreno et al., 2012).

Upper-plate variations may also affect nonvolcanic tremor occurrence. Along Cascadia, along-strike variations in tremor location and recurrence interval are spatially linked with variations in upper-plate geologic terranes (Brudzinski and Allen, 2007). Wells et al. (2017) showed a correlation between upper-plate faults and reduced levels of tremor activity along Cascadia, sug-gesting that the upper-plate faults allow fluid pressures to drain from the fault at depth, limiting tremor occurrence.

MEGATHRUST EARTHQUAKE HAZARDS

Due to their size and frequent occurrence, earthquakes on megathrust faults constitute significant seismic and tsunami hazards in subduction zones around the world. Slab deformation and partitioned block faulting of the upper plate also contribute to hazards in many regions.

Coseismic Slip and Geodetic Locking

With recent improvements in geodetic coverage by GPS and GNSS sta-tions on the upper plate of subduction zones, detailed images of hetero-geneous patterns of interseismic slip deficit (or “locking”) are now available.

Page 25: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1492Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

These illustrate the complexity of fault locking and creep in many subduction zones, and they provide a basis for refining the regional hazard. Comparisons between prior fault locking maps and recent coseismic slip in large events have led to a major advance, as relatively good correlations are observed between regions of high coseismic slip and strong plate locking when both are well resolved. For example, geodetic and paleogeodetic observations dat-ing back several decades suggest heterogeneous patterns of strong coupling along the Sumatra margin, and the strongly coupled zones served as the high-slip regions in the 2005 Mw 8.7 Nias, and 2007 Mw 8.4 and Mw 7.9 Sumatra earthquakes (Chlieh et al., 2008). Similar observations of strong locking found in regions of high coseismic slip exist for the 2007 Mw 8.0 Pisco, Peru (Perfettini et al., 2010), 2010 Mw 8.8 Maule, Chile (Moreno et al., 2010), 2011 Mw 9 Tohoku (e.g., Ozawa et al., 2012), 2012 Mw 7.8 Nicoya, Costa Rica (e.g., Protti et al., 2014), and the 2016 Mw 7.8 Ecuador (Ye et al., 2016a) earthquakes.

These correlations between the geodetically determined locked patches and areas of high slip during great earthquakes have implications for seis-mic and tsunami hazard assessments along subduction zones. However, land-based geodetic measurements have very poor resolution of near-trench slip deficit, as was found for the 2011 Tohoku earthquake (e.g., Lay, 2017). A major limitation is the paucity of offshore geodetic observations in most subduction zones; without offshore observations, resolution of the spatial extent of locking and strain accumulation is very limited. An example where offshore obser-vations now exist in reasonable density for this to be overcome is along the Nankai region of southwest Japan. Yokota et al. (2016) presented coupling esti-mates based on offshore measurements, suggesting wide areas of the shallow Nankai subduction zone have high slip deficit and are strongly locked, which expands the regions where shallow, tsunamigenic slip is possible. Seafloor measurements by GPS and acoustic methods (Sato et  al., 2011; Kido et  al., 2011), along with ocean bottom pressure sensors, provided invaluable data for resolving the shallow slip in the 2011 Tohoku earthquake. The greatly ex-panded cabled S-net offshore Honshu now provides continuous monitoring of moderate-slip earthquakes, and data assimilation methods will enable early tsunami warning of unprecedented quality.

Coseismic Slip in Aseismic Regions

The previous section outlined advances made in correlating areas of strong geodetic locking with the areas of high coseismic slip in great earthquakes; however, it is also valuable to evaluate low coupling zones. Areas thought to involve largely aseismic creep, in some cases linked to areas of rough sub-ducting topography (e.g., Wang and Bilek, 2014), can only be confirmed as deforming this way by geodetic measurements. We cannot ignore the pos-sibility than any subduction megathrust, including those with highly variable structure, can host large earthquakes, even if historical seismicity does not reveal the potential. One such example is along the Solomon Islands subduc-tion zone, where a triple junction marks the subduction of the Australia and

Solomon Islands–Woodlark Basin plates beneath the Pacific plate. This region was largely aseismic prior to the 2007 Mw 8.1 earthquake, which produced slip along the plate-boundary interface for both the Australian and Solomon Islands–Woodlark Basin plates, crossing the fracture zone boundary between these two plates and marking the first observed rupture across a triple junction (Taylor et al., 2008; Furlong et al., 2009).

The near-trench region of the megathrust was once considered to be an unlikely area to host large coseismic slip because of unfavorable frictional con-ditions, yet examples such as the massive coseismic slip in the near-trench region of the 2011 Tohoku earthquake (e.g., Lay et al., 2011b; Ide et al., 2011; Ozawa et al., 2012; Iinuma et al., 2012) provide dramatic evidence to the con-trary. Very shallow megathrust slip also occurred in the 2006 Mw 8.3 Kuril earth-quake, an area also lacking in previous interplate seismicity (Ammon et al., 2008; Lay et al., 2009).

Earthquake Triggering and Stress Transfer

Quantification of the stress transfer and earthquake triggering potential is another important consideration for assessing seismic hazard in subduction zones. Earthquake triggering by transfer of stress, either dynamic or static, is recognized in fault zones around the world (e.g., Hill et al., 1993; King et al., 1994; Lin and Stein, 2004; Freed, 2005). Many subduction zone examples exist, such as along the Sumatra margin, with its history of great earthquakes in the eighteenth and nineteenth centuries. Modern great earthquake activity in this region began with the 2004 Mw 9.2 Sumatra underthrusting event. Follow-ing the 2004 event, other great megathrust earthquake ruptures progressed southward along the margin in the 2005 Mw 8.6 Nias (e.g., Konca et al., 2007) and the 2007 Mw 8.5 and Mw 7.9 Sumatra earthquakes (e.g., Konca et al., 2008). This sequence brackets a locked region along the Mentawai Islands offshore of Padang, which last ruptured in 1797, and where a future great megathrust earthquake could occur.

The 2006–2007 Kuril Islands sequence is another example of coupled great earthquakes, although with diverse focal mechanisms. The first event in the sequence, the 2006 Mw 8.4 event, occurred on the megathrust, whereas the second Mw 8.1 event in 2007 was a normal-faulting, outer-rise event that produced stronger shaking in Japan due to higher moment-scaled radiated energy (e.g., Ammon et al., 2008). Slip from the 2006 earthquake, as well as from a large 1963 compressional earthquake in the trench slope region, led to increased static stress on optimally oriented normal faults in the region of the 2007 event (Raeesi and Atakan, 2009).

The 2009 Samoa-Tonga earthquake doublet also involved interaction between two separate faults, but with a pattern reversed in time relative to the 2006–2007 Kuril Islands sequence. In the Samoa-Tonga case, a Mw 8.1 normal- faulting, outer-rise event triggered rupture of the Tonga megathrust with cumu lative Mw of 8.0 within minutes, compounding tsunami runup and fatalities (Lay et al., 2010b). This style of triggering is particularly troublesome,

Page 26: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1493Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

because the secondary underthrusting events in the sequence occurred within tens of seconds, hampering the ability to quickly assess the tsunami potential.

The hazard posed by tsunami earthquakes is now well recognized, but it remains difficult to characterize in most regions. As mentioned herein, it is very challenging to establish whether the shallow portion of a megathrust is frictionally locked and accumulating strain using on-land geodetic data. This is particularly true for wide megathrusts. It is also difficult to establish the updip limit of rupture in large megathrust ruptures to judge whether there is potential for tsunami earthquakes to occur subsequently, as occurred in the 2010 Men-tawai earthquake updip from the 2007 Sumatra earthquakes. The tendency for very shallow ruptures to be depleted in short-period seismic wave radiation can cause delayed tsunami hazard reaction, even when shaking alerts coastal communities to an earthquake occurrence. The best strategy for mitigating this hazard appears to be offshore continuous seafloor deformation moni-toring (particularly cabled seafloor pressure sensors) and regional real-time source inversion procedures connected to an effective tsunami warning and evacuation system.

Strong shaking from offshore events constitutes another major hazard, and even less response time is available than for tsunami response. Continuous early warning systems that evaluate direct P-wave shaking with continuous up-dates may provide tens of seconds of warning time prior to strong S and sur-face-wave shaking, but the weak onset of the 2011 Tohoku earthquake indicates the challenges of evaluating how large an earthquake will grow after satura-tion of early high-sensitivity sensors. Continuous seismo-geodesy, combining strong motion sensors and real-time GPS, appears to be the most promising on-land method (e.g., Geng et al., 2013; Crowell et al., 2013; Melgar et al., 2013), with offshore cabled seafloor monitoring being even more valuable, although expensive.

In many regions, current instrumentation is deficient to achieve robust tsu-nami and seismic warning systems, and efforts to enhance seismic and geo-detic networks are needed. Societal preparation, enforced building codes, and seismic retrofitting of older structures are critical and essential elements for mitigating the hazard posed by megathrust earthquakes.

CONCLUSIONS

Megathrust faults on converging plate boundaries host Earth’s largest earthquakes, and the seismic and tsunami events in these regions constitute some of Earth’s primary natural hazards. Over the past 50 yr, great progress has been made in understanding the fundamental plate tectonics driving motions on megathrusts, but recent events such as the 2004 Sumatra and 2011 Japan events have still shocked even earthquake scientists with the devastating po-tential of megathrust faulting. Megathrust faults have diverse behavior that is partly understood in terms of overall tectonic setting, but questions remain concerning causes of frictional variations, possible loading rate dependence of frictional failure, the degree of shear stress reduction during faulting, and

even behavior of some regions with long repose between events. Advances in seismology and geodesy now provide information about the seismic cycle for some regions, and over time improved assessments of seismic potential appear to be viable. Offshore geodetic monitoring will provide critical infor-mation about the along-dip and near-trench strain accumulation process and the occurrence of shallow slow-slip events and stable sliding that may prevent strain buildup. Coupled offshore and onshore seismic and geodetic monitor-ing combined with real-time processing have the potential to establish early warning systems for megathrust tsunami and strong shaking, but this will re-quire extensive global investment in instrumentation and operations facilities around the world.

ACKNOWLEDGMENTS

T. Lay’s research on earthquakes is supported by National Science Foundation (NSF) grant EAR-1245717. S. Bilek’s research on earthquakes is supported by NSF grant OCE-1434550. Reviews by Guest Editor L. Wallace and two anonymous reviewers helped us to improve the manuscript.

REFERENCES CITED

Aki, K., 1965, Maximum likelihood estimate of b in the formula log n = a – bm and its confidence limits: Bulletin of the Earthquake Research Institute, Tokyo University, v. 43, p. 237–239.

Allmann, B.P., and Shearer, P.M., 2009, Global variations of stress drop for moderate to large earthquakes: Journal of Geophysical Research, v.  114, B01310, https:// doi .org /10 .1029 /2008JB005821 .

Ammon, C.J., Ji, C., Thio, H.K., Robinson, D., Ni, S.D., Hjorleifsdottir, V., Kanamori, H., Lay, T., Das, S., Helmberger, D., Ichinose, G., Polet, J., and Wald, D., 2005, Rupture process of the 2004 Sumatra-Andaman earthquake: Science, v.  308, p.  1133–1139, https:// doi .org /10 .1126 /science .1112260 .

Ammon, C.J., Kanamori, H., and Lay, T., 2008, A great earthquake doublet and seismic stress transfer cycle in the central Kuril Islands: Nature, v. 451, p. 561–565, https:// doi .org /10 .1038 /nature06521 .

Ando, M., 1975, Source mechanisms and tectonic significance of historical earthquakes along the Nankai Trough, Japan: Tectonophysics, v.  27, p.  119–140, https:// doi .org /10 .1016 /0040 -1951 (75)90102 -X .

Araki, E., Saffer, D.M., Kopf, A.J., Wallace, L.M., Kimura, T., Machida, Y., Ide, S., and Davis, E., 2017, Recurring and triggered slow-slip events near the trench at the Nankai Trough sub-duction megathrust: Science, v. 356, p. 1157–1160, https:// doi .org /10 .1126 /science .aan3120 .

Asano, Y., Obara, K., and Ito, Y., 2008, Spatiotemporal distribution of very-low frequency earth-quakes in Tokachi-oki near the junction of the Kuril and Japan trenches revealed by using array signal processing: Earth, Planets, and Space, v. 60, p. 871–875, https:// doi .org /10 .1186 /BF03352839 .

Asano, Y., Saito, T., Ito, Y., Shiomi, K., Hirose, H., Matsumoto, T., Aoi, S., Hori, S., and Sekiguchi, S., 2011, Spatial distribution and focal mechanisms of aftershocks of the 2011 off the Pacific coast of Tohoku earthquake: Earth, Planets, and Space, v. 63, p. 669–673, https:// doi .org /10 .5047 /eps .2011 .06 .016 .

Audet, P., and Kim, Y., 2016, Teleseismic constraints on the geological environment of deep episodic slow earthquakes in subduction zone forearcs: A review: Tectonophysics, v. 670, p. 1–15, https:// doi .org /10 .1016 /j .tecto .2016 .01 .005 .

Audet, P., and Schwartz, S.Y., 2013, Hydrologic control of forearc strength and seismicity in the Costa Rican subduction zone: Nature Geoscience, v. 6, p. 852–855, https:// doi .org /10 .1038 /ngeo1927 .

Audet, P., Bostock, M.G., Christensen, N.I., and Peacock, S.M., 2009, Seismic evidence for over-pressured subducted oceanic crust and megathrust fault sealing: Nature, v. 457, p. 76–78, https:// doi .org /10 .1038 /nature07650 .

Bangs, N.L., McIntosh, K.D., Silver, E.A., Kluesner, J.W., and Ranero, C.R., 2015, Fluid accumu-lation along the Costa Rica subduction thrust and development of the seismogenic zone:

Page 27: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1494Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

Journal of Geophysical Research–Solid Earth, v. 120, p. 67–86, https:// doi .org /10 .1002 /2014JB011265 .

Bassett, D., Sandwell, D.T., Fialko, Y., and Watts, A.B., 2016, Upper-plate controls on co-seismic slip in the 2011 magnitude 9.0 Tohoku-oki earthquake: Nature, v. 531, p. 92–96, https:// doi .org /10 .1038 /nature16945 .

Beavan, J., Wang, X., Holden, C., Wilson, K., Power, W., Prasetya, G., Bevis, M., and Kautoke, R., 2010, Near-simultaneous great earthquakes at Tongan megathrust and outer rise in Septem-ber 2009: Nature, v. 466, no. 7309, p. 959–963, https:// doi .org /10 .1038 /nature09292 .

Beck, S., Barrientos, S., Kausel, E., and Reyes, M., 1998, Source characteristics of historic earth-quakes along the central Chile subduction zone: Journal of South American Earth Sciences, v. 11, p. 115–129, https:// doi .org /10 .1016 /S0895 -9811 (98)00005 -4 .

Bell, R., Sutherland, R., Barker, D.H.N., Henrys, S., Bannister, S., Wallace, L., and Beavan, J., 2010, Seismic reflection character of the Hikurangi subduction interface, New Zealand, in the region of repeated Gisborne slow slip events: Geophysical Journal International, v. 180, p. 34–48, https:// doi .org /10 .1111 /j .1365 -246X .2009 .04401 .x .

Bell, R., Holden, C., Power, W., Wang, X.M., and Downes, G., 2014, Hikurangi margin tsunami earthquake generated by slow seismic rupture over a subducted seamount: Earth and Plan-etary Science Letters, v. 397, p. 1–9, https:// doi .org /10 .1016 /j .epsl .2014 .04 .005 .

Beroza, G.C., and Ide, S., 2011, Slow earthquakes and nonvolcanic tremor: Annual Review of Earth and Planetary Sciences, v. 39, p. 271–296, https:// doi .org /10 .1146 /annurev -earth -040809 -152531 .

Bilek, S.L., and Lay, T., 2002, Tsunami earthquakes possibly widespread manifestations of fric-tional conditional stability: Geophysical Research Letters, v. 29, 1673, https:// doi .org /10 .1029 /2002GL015215 .

Bilek, S.L., Schwartz, S.Y., and DeShon, H.R., 2003, Control of seafloor roughness on earthquake rupture behavior: Geology, v.  31, p.  455–458, https:// doi .org /10 .1130 /0091 -7613 (2003)031 <0455: COSROE>2 .0 .CO;2 .

Bilek, S.L., Engdahl, E.R., DeShon, H.R., and El Hariri, M., 2011, The 25 October 2010 Sumatra tsu-nami earthquake: Slip in a slow patch: Geophysical Research Letters, v. 38, L14306, https:// doi .org /10 .1029 /2011GL047864 .

Bilek, S.L., Rotman, H.M.M., and Phillips, W.S., 2016, Low stress drop earthquakes in the rup-ture zone of the 1992 Nicaragua tsunami earthquake: Geophysical Research Letters, v. 43, p. 10,180–10,188, https:// doi .org /10 .1002 /2016GL070409 .

Bletery, Q., Sladen, A., Delouis, B., Valée, M., Nocquet, J.-M., Rolland, L., and Jiang, J., 2014, A detailed source model for the MW 9.0 Tohoku-oki earthquake reconciling geodesy, seismol-ogy, and tsunami records: Journal of Geophysical Research–Solid Earth, v.  119, p.  7636–7653, https:// doi .org /10 .1002 /2014JB011261 .

Bletery, Q., Thomas, A.M., Rempel, A.W., Karlstrom, L., Sladen, A., and De Barros, L., 2016, Mega- earthquakes rupture flat megathrusts: Science, v.  354, p.  1027–1031, https:// doi .org /10 .1126 /science .aag0482 .

Bostock, M.G., Royer, A.A., Hearn, E.H., and Peacock, S.M., 2012, Low frequency earthquakes below southern Vancouver Island: Geochemistry Geophysics Geosystems, v.  13, Q11007, https:// doi .org /10 .1029 /2012GC004391 .

Briggs, R.W., Engelhart, S.E., Nelson, A.R., Dura, T., Kemp, A.C., Haeussler, P.J., Reide Corbett, D., Angster, S.J., and Bradly, L.-A., 2014, Uplift and subsidence reveal a nonpersistent mega thrust rupture boundary (Sitkinak Island, Alaska): Geophysical Research Letters, v. 41, p. 2289–2296, https:// doi .org /10 .1002 /2014GL059380 .

Brown, K.M., Tryon, M.D., DeShon, H.R., Dorman, L.M., and Schwartz, S.Y., 2005, Correlated transient fluid pulsing and seismic tremor in the Costa Rica subduction zone: Earth and Planetary Science Letters, v. 238, p. 189–203, https:// doi .org /10 .1016 /j .epsl .2005 .06 .055 .

Brown, J.R., Beroza, G.C., Ide, S., Ohta, K., Shelly, D.R., Schwartz, S.Y., Rabbel, W., Thorwart, M., and Kao, H., 2009, Deep low frequency earthquakes in tremor localize to the plate interface in multiple subduction zones: Geophysical Research Letters, v. 36, L19306, https:// doi .org /10 .1029 /2009GL040027 .

Brudzinski, M., and Allen, R.M., 2007, Segmentation in episodic tremor and slip all along Cas-cadia: Geology, v. 35, p. 907–910, https:// doi .org /10 .1130 /G23740A .1 .

Brudzinski, M., Cabral-Cano, E., Correa-Mora, F., DeMets, C., and Marquez-Azua, B.M., 2007, Slow slip transients along the Oaxaca subduction segment from 1993 to 2007: Geophysical Jour-nal International, v. 171, p. 523–538, https:// doi .org /10 .1111 /j .1365 -246X .2007 .03542 .x .

Carson, B., and Screaton, E.J., 1998, Fluid flow in accretionary prisms: Evidence for focused, time-variable discharge: Reviews of Geophysics, v.  36, p.  329–351, https:// doi .org /10 .1029 /97RG03633 .

Cesca, S., Grigoli, F., Heimann, S., Dahm, T., Kriegerowski, M., Sobiesiak, M., Tassara, C., and Olcay, M., 2016, The MW 8.1 2014 Iquique, Chile, seismic sequence: A tale of foreshocks and aftershocks: Geophysical Journal International, v. 204, p. 1766–1780, https:// doi .org /10 .1093 /gji /ggv544 .

Chen, T., Luo, H., and Furlong, K.P., 2017, A Bayesian rupture model of the 2007 MW 8.1 Solo-mon Islands earthquake in Southwest Pacific with coral reef displacement measurements: Journal of Asian Earth Sciences, v. 138, p. 92–97, https:// doi .org /10 .1016 /j .jseaes .2017 .01 .032 .

Chlieh, M., Avouac, J.P., Hjorleifsdottir, V., et al., 2007, Coseismic slip and afterslip of the great MW 9.15 Sumatra-Andaman earthquake of 2004: Bulletin of the Seismological Society of America, v. 97, p. S152–S173, https:// doi .org /10 .1785 /0120050631 .

Chlieh, M., Avouac, J.P., Sieh, K., Natawidjaja, D.H., and Galetzka, J., 2008, Heterogeneous coupling of the Sumatran megathrust constrained by geodetic and paleogeodetic mea-surements: Journal of Geophysical Research, v.  113, B05305, https:// doi .org /10 .1029 /2007JB004981 .

Choy, G.L., and Boatwright, J.L., 1995, Global patterns of radiated seismic energy and appar-ent stress: Journal of Geophysical Research, v. 100, p. 18,205–18,228, https:// doi .org /10 .1029 /95JB01969 .

Christensen, D.H., and Ruff, L.J., 1988, Seismic coupling and outer-rise earthquakes: Journal of Geophysical Research, v. 93, p. 13,421–13,444, https:// doi .org /10 .1029 /JB093iB11p13421 .

Cisternas, M., Carvajal, M., Wesson, R., Ely, L.L., and Gorigoitia, N., 2017, Exploring the historical earthquakes preceding the giant 1960 Chile earthquake in a time-dependent seismogenic zone: Bulletin of the Seismological Society of America, v. 107, p. 2664–2675, https:// doi .org /10 .1785 /0120170103 .

Cloos, M., 1992, Thrust-type subduction-zone earthquakes and seamount asperities—A physi-cal model for seismic rupture: Geology, v. 20, p. 601–604, https:// doi .org /10 .1130 /0091 -7613 (1992)020 <0601: TTSZEA>2 .3 .CO;2 .

Cloos, M., and Shreve, R.L., 1996, Shear-zone thickness and the seismicity of Chilean- and Mari-anas-type subduction zones: Geology, v.  24, p.  107–110, https:// doi .org /10 .1130 /0091 -7613 (1996)024 <0107: SZTATS>2 .3 .CO;2 .

Collot, J.Y., Marcaillou, B., Sage, F., Michaud, F., Agudelo, W., Charvis, P., Graindorge, D., Gutscher, M.A., and Spence, G., 2004, Are rupture zone limits of great subduction earth-quakes controlled by upper plate structures? Evidence from multichannel seismic reflection data acquired across the northern Ecuador–southwest Colombia margin: Journal of Geo-physical Research, v. 109, B11103, https:// doi .org /10 .1029 /2004JB003060 .

Convers, J.A., and Newman, A.V., 2011, Global evaluation of large earthquake energy from 1997 through mid-2010: Journal of Geophysical Research, v. 116, B08304, https:// doi .org /10 .1029 /2010JB007928 .

Correa-Mora, F., DeMets, C., Cabral-Cano, E., Diaz-Molina, O., and Marquez-Azua, B., 2009, Transient deformation in southern Mexico in 2006 and 2007: Evidence for distinct deep-slip patches beneath Guerrero and Oaxaca: Geochemistry Geophysics Geosystems, v. 10, Q02S12, https:// doi .org /10 .1029 /2008GC002211 .

Crowell, B.W., Melgar, D., Bock, Y., Haase, J.S., and Geng, J.H., 2013, Earthquake magnitude scal-ing using seismogeodetic data: Geophysical Research Letters, v. 40, p. 6089–6094, https:// doi .org /10 .1002 /2013GL058391 .

Das, S., and Henry, C., 2003, Spatial relation between main earthquake slip and its aftershock distribution: Reviews of Geophysics, v. 41, 1013, https:// doi .org /10 .1029 /2002RG000119 .

den Hartog, S.A.M., and Spiers, C.J., 2013, Influence of subduction zone conditions and gouge composition on frictional slip stability of megathrust faults: Tectonophysics, v. 600, p. 75–90, https:// doi .org /10 .1016 /j .tecto .2012 .11 .006 .

den Hartog, S.A.M., Niemeijer, A.R., and Spiers, C.J., 2012, New constraints on megathrust slip stability under subduction zone P-T conditions: Earth and Planetary Science Letters, v. 353, p. 240–252, https:// doi .org /10 .1016 /j .epsl .2012 .08 .022 .

Denolle, M.A., and Shearer, P.M., 2016, New perspectives on self-similarity for shallow thrust earthquakes: Journal of Geophysical Research, v. 121, p. 6533–6565, https:// doi .org /10 .1002 /2016JB013105 .

Ding, M., and Lin, J., 2016, Deformation and faulting of subduction overriding plate caused by a subducted seamount: Geophysical Research Letters, v. 43, p. 8936–8944, https:// doi .org /10 .1002 /2016GL069785 .

Dominguez, S., Malavielle, J., and Lallemand, S.E., 2000, Deformation of accretionary wedges in response to seamount subduction: Insight from sandbox experiments: Tectonics, v. 19, p. 182–196, https:// doi .org /10 .1029 /1999TC900055 .

Page 28: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1495Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

Dragert, H., Wang, K.L., and James, T.S., 2001, A silent slip event on the deeper Cascadia subduc-tion interface: Science, v. 292, p. 1525–1528, https:// doi .org /10 .1126 /science .1060152 .

Duan, B., 2012, Dynamic rupture of the 2011 Mw 9.0 Tohoku-oki earthquake: Roles of a possible subducting seamount: Journal of Geophysical Research, v.  117, B05311, https:// doi .org /10 .1029 /2011JB009124 .

Eberhart-Phillips, D., and Bannister, S., 2015, 3-D imaging of the northern Hikurangi subduction zone, New Zealand: Variations in subducted sediment, slab fluids and slow slip: Geophysical Journal International, v. 201, p. 838–855, https:// doi .org /10 .1093 /gji /ggv057 .

Ekström, G., Nettles, M., and Dziewonski, A.M., 2012, The global CMT project 2004-2010: Cen-troid-moment tensors for 13,017 earthquakes: Physics of the Earth and Planetary Interiors, v. 200, p. 1–9, https:// doi .org /10 .1016 /j .pepi .2012 .04 .002 .

El Hariri, M., and Bilek, S.L., 2011, Stress changes and aftershock distribution of the 1994 and 2006 Java subduction zone earthquake sequences: Journal of Geophysical Research, v. 116, B06306, https:// doi .org /10 .1029 /2010JB008124 .

El Hariri, M., Bilek, S.L., DeShon, H.R., Engdahl, E.R., and Bisrat, S., 2013, Along-strike variabil-ity of rupture duration in subduction zone earthquakes: Journal of Geophysical Research, v. 118, p. 646–664, https:// doi .org /10 .1029 /2012jb009548 .

Fan, W., Shearer, P.M., Ji, C., and Bassett, D., 2016, Multiple branching rupture of the 2009 Tonga- Samoa earthquake: Journal of Geophysical Research–Solid Earth, v.  121, p.  5809–5827, https:// doi .org /10 .1002 /2016JB012945 .

Farías, M., Comte, D., Roecker, S., Carrizo, D., and Pardo, M., 2011, Crustal extensional faulting triggered by the 2010 Chilean earthquake: The Pichilemu seismic sequence: Tectonics, v. 30, TC6010, https:// doi .org /10 .1029 /2011TC002888 .

Frank, W.B., Shapiro, N.M., Husker, A.L., Kostoglodov, V., Bhat, H.S., and Campillo, M., 2015, Along-fault pore-pressure evolution during a slow-slip event in Guerrero, Mexico: Earth and Planetary Science Letters, v. 413, p. 135–143, https:// doi .org /10 .1016 /j .epsl .2014 .12 .051 .

Freed, A.M., 2005, Earthquake triggering by static, dynamic, and postseismic stress transfer: Annual Review of Earth and Planetary Sciences, v.  33, p.  335–367, https:// doi .org /10 .1146 /annurev .earth .33 .092203 .122505 .

Frohlich, C., and Davis, S.D., 1993, Teleseismic b values; Or, much ado about 1.0: Journal of Geo-physical Research, v. 98, p. 631–644, https:// doi .org /10 .1029 /92JB01891 .

Fujii, Y., and Satake, K., 2007, Tsunami source of the 2004 Sumatra-Andaman earthquake inferred from tide gauge and satellite data: Bulletin of the Seismological Society of America, v. 97, p. S192–S207, https:// doi .org /10 .1785 /0120050613 .

Fujiwara, T., Kodaira, S., No, T., Kaiho, Y., and Kaneda, Y., 2011, The 2011 Tohoku-oki earthquake: Displacement reaching the trench axis: Science, v.  334, p.  1240, https:// doi .org /10 .1126 /science .1211554 .

Furlong, K.P., Lay, T., and Ammon, C.J., 2009, A great earthquake rupture across a rapidly evolving three-plate boundary: Science, v.  324, p.  226–229, https:// doi .org /10 .1126 /science .1167476 .

Geersen, J., McNeill, L., Henstock, T.J., and Gaedicke, C., 2013, The 2004 Aceh-Andaman earth-quake: Early clay dehydration controls shallow seismic rupture: Geochemistry Geophysics Geosystems, v. 14, p. 3315–3323, https:// doi .org /10 .1002 /ggge .20193 .

Geersen, J., Ranero, C.R., Barckhausen, U., and Reichert, C., 2015, Subducting seamounts con-trol interplate coupling and seismic rupture in the 2014 Iquique earthquake area: Nature Communications, v. 6, 8267, https:// doi .org /10 .1038 /ncomms9267 .

Geng, J.H., Bock, Y., Melgar, D., Crowell, B.W., and Haase, J.S., 2013, A new seismogeodetic ap-proach applied to GPS and accelerometer observations of the 2012 Brawley seismic swarm: Implications for earthquake early warning: Geochemistry Geophysics Geosystems, v.  14, p. 2124–2142, https:// doi .org /10 .1002 /ggge .20144 .

Ghosh, A., Newman, A.V., Thomas, A.M., and Farmer, G.T., 2008, Interface locking along the sub-duction megathrust from b-value mapping near Nicoya Peninsula, Costa Rica: Geophysical Research Letters, v. 35, L01301, https:// doi .org /10 .1029 /2007GL031617 .

Grevemeyer, I., Ranero, C.R., and Ivandic, M., 2018, Structure of oceanic crust and serpentini-zation at subduction trenches: Geosphere, v.  14, p.  395–418, https:// doi .org /10 .1130 /GES01537 .1 .

Gutenberg, B., and Richter, C.F., 1944, Frequency of earthquakes in California: Bulletin of the Seismological Society of America, v. 34, p. 185–188.

Hayes, G.P., 2017, The finite, kinematic rupture properties of great-sized earthquakes since 1990: Earth and Planetary Science Letters, v. 468, p. 94–100, https:// doi .org /10 .1016 /j .epsl .2017 .04 .003 .

Hayes, G.P., Furlong, K.P., Benz, H.M., and Herman, M.W., 2014a, Triggered aseismic slip adjacent to the 6 February 2013 MW 8.0 Santa Cruz Islands megathrust earthquake: Earth and Plane-tary Science Letters, v. 388, p. 265–272, https:// doi .org /10 .1016 /j .epsl .2013 .11 .010 .

Hayes, G.P., Herman, M.W., Barnhart, W.D., Furlong, K.P., Riquelme, S., Benz, H.M., Bergman, E., Barrientos, S., Earle, P.S., and Samsonov, S., 2014b, Continuing megathrust earthquake potential in Chile after the 2014 Iquique earthquake: Nature, v. 512, p. 295–298, https:// doi .org /10 .1038 /nature13677 .

Henstock, T.J., McNeill, L.C., Bull, J.M., Cook, B.J., Gulick, S.P.S., Austin, J.A., Permana, H., and Djajadihardja, Y.S., 2016, Downgoing plate topography stopped rupture in the A.D. 2005 Sumatra earthquake: Geology, v. 44, p. 71–74, https:// doi .org /10 .1130 /G37258 .1 .

Heuret, A., Conrad, C.P., Funiciello, F., Lallemand, S., and Sandri, L., 2012, Relation between sub-duction megathrust earthquakes, trench sediment thickness and upper plate strain: Geo-physical Research Letters, v. 39, L05304, https:// doi .org /10 .1029 /2011GL050712 .

Hicks, S.P., and Rietbrock, A., 2015, Seismic slip on an upper-plate normal fault during a large subduction megathrust rupture: Nature Geoscience, v. 8, p. 955–960, https:// doi .org /10 .1038 /ngeo2585 .

Hicks, S.P., Rietbrock, A., Haberland, C.A., Ryder, I.M.A., Simons, M., and Tassara, A., 2012, The 2010 MW 8.8 Maule, Chile, earthquake: Nucleation and rupture propagation controlled by a subducted topographic high: Geophysical Research Letters, v. 39, L19308, https:// doi .org /10 .1029 /2012GL053184 .

Hill, D.P., Reasenberg, P.A., Michael, A., et al., 1993, Seismicity remotely triggered by the mag-nitude 7.3 Landers, California, earthquake: Science, v. 260, p. 1617–1623, https:// doi .org /10 .1126 /science .260 .5114 .1617 .

Hirose, H., and Obara, K., 2006, Short-term slow slip and correlated tremor episodes in the Tokai region, central Japan: Geophysical Research Letters, v.  33, L17311, https:// doi .org /10 .1029 /2006GL026579 .

Holtkamp, S.G., and Brudzinski, M.R., 2011, Earthquake swarms in circum-Pacific subduction zones: Earth and Planetary Science Letters, v. 305, p. 215–225, https:// doi .org /10 .1016 /j .epsl .2011 .03 .004 .

Hsu, Y.J., Simons, M., Avouac, J.P., Galetzka, J., Sieh, K., Chlieh, M., Natawidjaja, D., Prawiro-dirdjo, L., and Bock, Y., 2006, Frictional afterslip following the 2005 Nias-Simeulue earth-quake, Sumatra: Science, v. 312, p. 1921–1926, https:// doi .org /10 .1126 /science .1126960 .

Huang, Z., and Zhao, D., 2013, Mechanism of the 2011 Tohoku-oki earthquake (Mw 9.0) and tsu-nami: Insight from seismic tomography: Journal of Asian Earth Sciences, v. 70–71, p. 160–168, https:// doi .org /10 .1016 /j .jseaes .2013 .03 .010 .

Husen, S., Kissling, E., and Quintero, R., 2002, Tomographic evidence for a subducted sea-mount beneath the Gulf of Nicoya, Costa Rica: The cause of the 1990 MW  = 7.0 Gulf of Nicoya earthquake: Geophysical Research Letters, v.  29,  1238, https:// doi .org /10 .1029 /2001GL014045 .

Ide, S., 2010, Striations, duration, migration and tidal response in deep tremor: Nature, v. 466, p. 356–359, https:// doi .org /10 .1038 /nature09251 .

Ide, S., Beroza, G.C., Shelly, D.R., and Uchide, T., 2007a, A scaling law for slow earthquakes: Nature, v. 447, p. 76–79, https:// doi .org /10 .1038 /nature05780 .

Ide, S., Shelly, D.R., and Beroza, G.C., 2007b, Mechanism of deep low frequency earthquakes: Further evidence that deep non-volcanic tremor is generated by shear slip on the plate inter-face: Geophysical Research Letters, v. 34, L03308, https:// doi .org /10 .1029 /2006GL028890 .

Ide, S., Baltay, A., and Beroza, G.C., 2011, Shallow dynamic overshoot and energetic deep rup-ture in the 2011 MW 9.0 Tohoku-oki earthquake: Science, v. 332, p. 1426–1429, https:// doi .org /10 .1126 /science .1207020 .

Ihmlé, P.F., 1996a, Frequency-dependent relocation of the 1992 Nicaragua slow earthquake: An empirical Green’s function approach: Geophysical Journal International, v.  127, p.  75–85, https:// doi .org /10 .1111 /j .1365 -246X .1996 .tb01536 .x .

Ihmlé, P.F., 1996b, Monte Carlo slip inversion in the frequency domain: Application to the 1992 Nicaragua slow earthquake: Geophysical Research Letters, v. 23, p. 913–916, https:// doi .org /10 .1029 /96GL00872 .

Iinuma, T., Hino, R., Kido, M., et al., 2012, Coseismic slip distribution of the 2011 off the Pacific coast of Tohoku earthquake (M 9.0) refined by means of seafloor geodetic data: Journal of Geophysical Research, v. 117, B07409, https:// doi .org /10 .1029 /2012JB009186 .

Ikari, M.J., Hüpers, A., and Kopf, A.J., 2013, Shear strength of sediments approaching subduc-tion in the Nankai Trough, Japan, as constraints on forearc mechanics: Geochemistry Geo-physics Geosystems, v. 14, p. 2716–2730, https:// doi .org /10 .1002 /ggge .20156 .

Page 29: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1496Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

Ikari, M.J., Ito, Y., Ujiie, K., and Kopf, A.J., 2015a, Spectrum of slip behaviour in Tohoku fault zone samples at plate tectonic slip rates: Nature Geoscience, v. 8, p. 870–874, https:// doi .org /10 .1038 /ngeo2547 .

Ikari, M.J., Kameda, J., Saffer, D.M., and Kopf, A.J., 2015b, Strength characteristics of Japan Trench borehole samples in the high-slip region of the 2011 Tohoku-oki earthquake: Earth and Planetary Science Letters, v. 412, p. 35–41, https:// doi .org /10 .1016 /j .epsl .2014 .12 .014 .

Ishii, M., 2011, High-frequency rupture properties of the Mw 9.0 off the Pacific coast of Tohoku earthquake: Earth, Planets, and Space, v. 63, p. 609–614, https:// doi .org /10 .5047 /eps .2011 .07 .009 .

Ito, Y., and Obara, K., 2006, Very low frequency earthquakes within accretionary prisms are very low stress-drop earthquakes: Geophysical Research Letters, v. 33, L09302, https:// doi .org /10 .1029 /2006GL025883.

Ito, Y., Obara, K., Shiomi, K., Sekine, S., and Hirose, H., 2007, Slow earthquakes coincident with episodic tremors and slow slip events: Science, v. 315, p. 503–506, https:// doi .org /10 .1126 /science .1134454 .

Ito, Y., Obara, K., Matsuzawa, T., and Maeda, T., 2009, Very low frequency earthquakes related to small asperities on the plate boundary interface at the locked to aseismic transition: Journal of Geophysical Research, v. 114, B00A13, https:// doi .org /10 .1029 /2008JB006036 .

Ito, Y., Hino, R., Kido, M., et al., 2013, Episodic slow slip events in the Japan subduction zone before the 2011 Tohoku-oki earthquake: Tectonophysics, v. 600, p. 14–26, https:// doi .org /10 .1016 /j .tecto .2012 .08 .022 .

Jiang, Y., Wdowinski, S., Dixon, T.H., Hackl, M., Protti, M., and Gonzalez, V., 2012, Slow slip events in Costa Rica detected by continuous GPS observations, 2002–2011: Geochemistry Geo-physics Geosystems, v. 13, Q04006, https:// doi .org /10 .1029 /2012GC004058 .

Kagan, Y.Y., and Jackson, D.D., 1995, New seismic gap hypothesis: Five years after: Journal of Geophysical Research, v. 100, p. 3943–3959, https:// doi .org /10 .1029 /94JB03014 .

Kagan, Y.Y., Jackson, D.D., and Geller, R.J., 2012, Characteristic earthquake model, 1884–2011, R.I.P.: Seismological Research Letters, v. 83, p. 951–953, https:// doi .org /10 .1785 /0220120107 .

Kanamori, H., 1971, Seismological evidence for a lithospheric normal faulting—The Sanriku earthquake of 1933: Physics of the Earth and Planetary Interiors, v. 4, p. 289–300, https:// doi .org /10 .1016 /0031 -9201 (71)90013 -6 .

Kanamori, H., 1972, Mechanism of tsunami earthquakes: Physics of the Earth and Planetary Inte-riors, v. 6, p. 346–359, https:// doi .org /10 .1016 /0031 -9201 (72)90058 -1 .

Kanamori, H., 2014, The diversity of large earthquakes and its implications for hazard mitiga-tion: Annual Review of Earth and Planetary Sciences, v. 42, p. 7–26, https:// doi .org /10 .1146 /annurev -earth -060313 -055034 .

Kanamori, H., and Brodsky, E.E., 2004, The physics of earthquakes: Reports on Progress in Physics, v. 67, p. 1429–1496, https:// doi .org /10 .1088 /0034 -4885 /67 /8 /R03 .

Kanamori, H., and Kikuchi, M., 1993, The 1992 Nicaragua earthquake: A slow tsunami earth-quake associated with subducted sediments: Nature, v. 361, p. 714–716, https:// doi .org /10 .1038 /361714a0 .

Kanamori, H., and McNally, K.C., 1982, Variable rupture mode of the subduction zone along the Ecuador-Colombia coast: Bulletin of the Seismological Society of America, v. 72, no. 4, p. 1241–1253.

Kanamori, H., Rivera, L., and Lee, W.H.K., 2010, Historical seismograms for unravelling a myste-rious earthquake: The 1907 Sumatra earthquake: Geophysical Journal International, v. 183, p. 358–374, https:// doi .org /10 .1111 /j .1365 -246X .2010 .04731 .x .

Kato, A., and Nakagawa, S., 2014, Multiple slow-slip events during a foreshock sequence of the 2014 Iquique, Chile, MW 8.1 earthquake: Geophysical Research Letters, v. 41, p. 5420–5427, https:// doi .org /10 .1002 /2014GL061138 .

Kato, A., Iidaka, T., Ikuta, R., et al., 2010, Variations of fluid pressure within the subducting oceanic crust and slow earthquakes: Geophysical Research Letters, v. 37, L14310, https:// doi .org /10 .1029 /2010GL043723 .

Kato, A., Obara, K., Igarashi, T., Tsuruoka, H., Nakagawa, S., and Hirata, N., 2012, Propagation of slow slip leading up to the 2011 MW 9.0 Tohoku-oki earthquake: Science, v. 335, p. 705–708, https:// doi .org /10 .1126 /science .1215141 .

Kato, A., Sakai, S., and Obara, K., 2011, A normal-faulting seismic sequence triggered by the 2011 off the Pacific coast of Tohoku earthquake: Wholesale stress regime changes in the upper plate: Earth, Planets, and Space, v. 63, p. 43, https:// doi .org /10 .5047 /eps .2011 .06 .014 .

Katsumata, A., and Kamaya, N., 2003, Low-frequency continuous tremor around the Moho discontinuity away from volcanoes in the southwest Japan: Geophysical Research Letters, v. 30, 1020, https:// doi .org /10 .1029 /2002GL015981 .

Kelleher, J.A., 1972, Rupture zones of large South American earthquakes and some predic-tions: Journal of Geophysical Research, v. 77, no. 11, p. 2087–2103, https:// doi .org /10 .1029 /JB077i011p02087 .

Kelleher, J.A., and McCann, W., 1976, Buoyant zones, great earthquakes, and unstable bound-aries of subduction: Journal of Geophysical Research, v. 81, p. 4885–4896, https:// doi .org /10 .1029 /JB081i026p04885 .

Kido, M., Osada, Y., Fujimoto, H., Hino, R., and Ito, Y., 2011, Trench-normal variation in observed seafloor displacements associated with the 2011 Tohoku-oki earthquake: Geophysical Re-search Letters, v. 38, L24303, https:// doi .org /10 .1029 /2011GL050057 .

Kikuchi, M., and Kanamori, H., 1995, Source characteristics of the 1992 Nicaragua tsunami earth-quake inferred from teleseismic body waves: Pure and Applied Geophysics, v. 144, p. 441–453, https:// doi .org /10 .1007 /BF00874377 .

King, G.C.P., Stein, R.S., and Lin, J., 1994, Static stress changes and the triggering of earthquakes: Bulletin of the Seismological Society of America, v. 84, p. 935–953.

Kiser, E., and Ishii, M., 2011, The 2010 MW 8.8 Chile earthquake: Triggering on multiple segments and frequency-dependent rupture behavior: Geophysical Research Letters, v.  38, L07301, https:// doi .org /10 .1029 /2011GL047140 .

Klein, E., Vigny, C., Fleitout, L., Grandin, R., Jolivet, R., Rivera, E., and Metois, M., 2017, A com-prehensive analysis of the Illapel 2015 MW 8.3 earthquake from GPS and InSAR data: Earth and Planetary Science Letters, v. 469, p. 123–134, https:// doi .org /10 .1016 /j .epsl .2017 .04 .010 .

Kodaira, S., Takahashi, N., Nakanishi, A., Miura, S., and Kaneda, Y., 2000, Subducted seamount imaged in the rupture zone of the 1946 Nankaido earthquake: Science, v. 289, p. 104–106, https:// doi .org /10 .1126 /science .289 .5476 .104 .

Kodaira, S., Iidaka, T., Kato, A., Park, J.O., Iwasaki, T., and Kaneda, Y., 2004, High pore fluid pres-sure may cause silent slip in the Nankai Trough: Science, v. 304, p. 1295–1298, https:// doi .org /10 .1126 /science .1096535 .

Konca, A.O., Hjorleifsdottir, V., Song, T.R.A., Avouac, J.P., Helmberger, D.V., Ji, C., Sieh, K., Briggs, R., and Meltzner, A., 2007, Rupture kinematics of the 2005 Mw 8.6 Nias-Simeulue earthquake from the joint inversion of seismic and geodetic data: Bulletin of the Seismological Society of America, v. 97, p. S307–S322, https:// doi .org /10 .1785 /0120050632 .

Konca, A.O., Avouac, J-P., Sladen, A., et al., 2008, Partial rupture of a locked patch of the Sumatra megathrust during the 2007 earthquake sequence: Nature, v. 456, p. 631–635, https:// doi .org /10 .1038 /nature07572 .

Koper, K.D., Hutko, A.R., Lay, T., Ammon, C.J., and Kanamori, H., 2011, Frequency-dependent rupture process of the 2011 MW 9.0 Tohoku earthquake: Comparison of short-period P wave backprojection images and broadband seismic rupture models: Earth, Planets, and Space, v. 63, p. 16, https:// doi .org /10 .5047 /eps .2011 .05 .026 .

Koper, K.D., Hutko, A.R., Lay, T., and Sufri, O., 2012, Imaging short-period seismic radiation from the 27 February 2010 Chile (MW 8.8) earthquake by back-projection of P, PP, and PKIKP waves: Journal of Geophysical Research, v. 117, B02308, https:// doi .org /10 .1029 /2011JB008576 .

Korenaga, J., 2017, On the extent of mantle hydration caused by plate bending: Earth and Plane-tary Science Letters, v. 457, p. 1–9, https:// doi .org /10 .1016 /j .epsl .2016 .10 .011 .

Kostoglodov, V., Singh, S.K., Santiago, J.A., Franco, S.I., Larson, K.M., Lowry, A.R., and Bilham, R., 2003, A large silent earthquake in the Guerrero seismic gap, Mexico: Geophysical Re-search Letters, v. 30, 1807, https:// doi .org /10 .1029 /2003GL017219 .

Kurzawski, R.M., Stipp, M., Niemeijer, A.R., Spiers, C.J., and Behrmann, J.H., 2016, Earthquake nucleation in weak subducted carbonates: Nature Geoscience, v. 9, p. 717–722, https:// doi .org /10 .1038 /ngeo2774 .

La Femina, P., Dixon, T.H., Govers, R., Norabuena, E., Turner, H., Saballos, A., Mattioli, G., Protti, M., and Strauch, W., 2009, Fore-arc motion and Cocos Ridge collision in Central America: Geochemistry Geophysics Geosystems, v.  10, Q05514, https:// doi .org /10 .1029 /2008GC002181 .

Lallemand, S., Heuret, A., and Boutelier, D., 2005, On the relationships between slab dip, back-arc stress, upper plate absolute motion, and crustal nature in subduction zones: Geochemis-try Geophysics Geosystems, v. 6, Q09006, https:// doi .org /10 .1029 /2005GC000917 .

Lange, D., Tilmann, F., Barrientos, S.E., Contreras-Reyes, E., Methe, P., Moreno, M., Heit, B., Agurto, H., Bernard, P., Vilotte, J.-P., and Beck, S., 2012, Aftershock seismicity of the 27 Febru-ary 2010 MW 8.8 Maule earthquake rupture zone: Earth and Planetary Science Letters, v. 317, p. 413–425, https:// doi .org /10 .1016 /j .epsl .2011 .11 .034 .

Lay, T., 2015, The surge of great earthquakes from 2004 to 2014: Earth and Planetary Science Letters, v. 409, p. 133–146, https:// doi .org /10 .1016 /j .epsl .2014 .10 .047 .

Page 30: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1497Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

Lay, T., 2017, A review of the rupture characteristics of the 2011 Tohoku-oki MW 9.1 earthquake: Tectonophysics, v. 733, p. 4–36, https:// doi .org /10 .1016 /j .tecto .2017 .09 .022 .

Lay, T., and Bilek, S.L., 2007, Anomalous earthquake ruptures at shallow depths on subduction zone megathrusts, in Dixon, T.H., and Moore, J.C., eds., The Seismogenic Zone of Subduc-tion Thrust Faults: New York, Columbia University Press, p. 476–511, https:// doi .org /10 .7312 /dixo13866 -015.

Lay, T., and Kanamori, H., 2011, Insights from the great 2011 Japan earthquake: Physics Today, v. 64, no. 12, p. 33–39, https:// doi .org /10 .1063 /PT .3 .1361 .

Lay, T., Kanamori, H., and Ruff, L., 1982, The asperity model and the nature of large subduction zone earthquake occurrence: Earthquake Prediction Research, v. 1, p. 3–71.

Lay, T., Kanamori, H., Ammon, C.J., et al., 2005, The great Sumatra-Andaman earthquake of 26 December 2004: Science, v. 308, p. 1127–1133, https:// doi .org /10 .1126 /science .1112250 .

Lay, T., Kanamori, H., Ammon, C.J., Hutko, A.R., Furlong, K., and Rivera, L., 2009, The 2006–2007 Kuril Islands great earthquake sequence: Journal of Geophysical Research, v. 114, B11308, https:// doi .org /10 .1029 /2008JB006280 .

Lay, T., Ammon, C.J., Hutko, A.R., and Kanamori, H., 2010a, Effects of kinematic constraints on teleseismic finite-source rupture inversions: Great Peruvian earthquakes of 23 June 2011 and 15 August 2007: Bulletin of the Seismological Society of America, v. 100, p. 969–994, https:// doi .org /10 .1785 /0120090274 .

Lay, T., Ammon, C.J., Kanamori, H., Rivera, L., Koper, K.D., and Hutko, A.R., 2010b, The 2009 Samoa-Tonga great earthquake triggered doublet: Nature, v. 466, p. 964–968, https:// doi .org /10 .1038 /nature09214 .

Lay, T., Ammon, C.J., Kanamori, H., Yamazaki, Y., Cheung, K.F., and Hutko, A.R., 2011a, The 25 October 2010 Mentawai tsunami earthquake (Mw 7.8) and the tsunami hazard presented by shallow megathrust ruptures: Geophysical Research Letters, v. 38, L06302, https:// doi .org /10 .1029 /2010GL046552 .

Lay, T., Ammon, C.J., Kanamori, H., Xue, L., and Kim, M.J., 2011b, Possible large near-trench slip during the 2011 MW 9.0 off the Pacific coast of Tohoku earthquake: Earth, Planets, and Space, v. 63, p. 687–692, https:// doi .org /10 .5047 /eps .2011 .05 .033 .

Lay, T., Kanamori, H., Ammon, C.J., Koper, K.D., Hutko, A.R., Ye, L.L., Yue, H., and Rushing, T.M., 2012, Depth-varying rupture properties of subduction zone megathrust faults: Journal of Geophysical Research, v. 117, B04311, https:// doi .org /10 .1029 /2011JB009133 .

Lay, T., Ye, L., Kanamori, H., Yamazaki, Y., Cheung, K.F., and Ammon, C.J., 2013, The February 6, 2013, MW 8.0 Santa Cruz Islands earthquake and tsunami: Tectonophysics, v. 608, p. 1109–1121, https:// doi .org /10 .1016 /j .tecto .2013 .07 .001 .

Lay, T., Yue, H., Brodsky, E.E., and An, C., 2014, The 1 April 2014 Iquique, Chile, MW 8.1 earthquake rupture sequence: Geophysical Research Letters, v. 41, p. 3818–3825, https:// doi .org /10 .1002 /2014GL060238 .

Lay, T., Ye, L., Ammon, C.J., and Kanamori, H., 2017, Intraslab rupture triggering megathrust rup-ture coseismically in the 17 December 2016 Solomon Islands MW 7.9 earthquake: Geophysi-cal Research Letters, v. 44, p. 1286–1292, https:// doi .org /10 .1002 /2017GL072539 .

Li, L., Lay, T., Cheung, K.F., and Ye, L., 2016, Joint modeling of teleseismic and tsunami wave observations to constrain the 16 September 2015 Illapel, Chile, MW 8.3 earthquake rup-ture process: Geophysical Research Letters, v.  43, p.  4303–4312, https:// doi .org /10 .1002 /2016GL068674 .

Lin, J., and Stein, R.S., 2004, Stress triggering in thrust and subduction earthquakes and stress interaction between the southern San Andreas and nearby thrust and strike-slip faults: Jour-nal of Geophysical Research, v. 109, B02303, https:// doi .org /10 .1029 /2003JB002607 .

Liu, C., Zheng, Y., Xiong, X., Wang, R., Lopez, A., and Li, J., 2015, Rupture process of the 2012 September 5 MW 7.6 Nicoya, Costa Rica, earthquake constrained by improved geodetic and seismological observations: Geophysical Journal International, v. 203, p. 175–183, https:// doi .org /10 .1093 /gji /ggv295 .

Lotto, G.C., Dunham, E.M., Jeppson, T.N., and Tobin, H.J., 2017, The effect of compliant prisms on subduction zone earthquakes and tsunamis: Earth and Planetary Science Letters, v. 458, p. 213–222, https:// doi .org /10 .1016 /j .epsl .2016 .10 .050 .

Maksymowicz, A., Chadwell, C.D., Ruiz, J., Tréhu, A.M., Contreras-Reyes, E., Weinrebe, W., Díaz-Naveas, J., Gibson, J.C., Lonsdale, P., and Tryon, M.D., 2017, Coseismic seafloor deformation in the trench region during the Mw 8.8 Maule megathrust earthquake: Scientific Reports, v. 7, p. 45918, https:// doi .org /10 .1038 /srep45918 .

Marcaillou, B., Collot, J.Y., Ribodetti, A., d’Acremont, E., Mahamat, A.A., and Alvarado, A., 2016, Seamount subduction at the north Ecuadorian convergent margin: Effects on structures,

inter-seismic coupling and seismogenesis: Earth and Planetary Science Letters, v.  433, p. 146–158, https:// doi .org /10 .1016 /j .epsl .2015 .10 .043 .

Matsuzawa, T., Asano, Y., and Obara, K., 2015, Very low frequency earthquakes off the Pacific coast of Tohoku, Japan: Geophysical Research Letters, v. 42, p. 4318–4325, https:// doi .org /10 .1002 /2015GL063959 .

McCaffrey, R., Long, M.D., Goldfinger, C., Zwick, P.C., Nabelek, J.L., Johnson, C.K., and Smith, C., 2000, Rotation and plate locking at the southern Cascadia subduction zone: Geophysical Research Letters, v. 27, p. 3117–3120, https:// doi .org /10 .1029 /2000GL011768 .

McCann, W.R., Nishenko, S.P., Sykes, L.R., and Kraus, J., 1979, Seismic gaps and plate tectonics: Seismic potential for major boundaries: Pure and Applied Geophysics, v. 117, p. 1082–1147, https:// doi .org /10 .1007 /BF00876211 .

Meier, M.-A., Ampuero, J.P., and Heaton, T., 2017, The hidden simplicity of subduction mega-thrust earthquakes: Science, v. 357, p. 1277–1281, https:// doi .org /10 .1126 /science .aan5643 .

Melgar, D., and Bock, Y., 2015, Kinematic earthquake source inversion and tsunami runup pre-diction with regional geophysical data: Journal of Geophysical Research–Solid Earth, v. 120, p. 3324–3349, https:// doi .org /10 .1002 /2014JB011832 .

Melgar, D., and Hayes, G.P., 2017, Systematic observations of the slip pulse properties of large earthquake ruptures: Geophysical Research Letters, v. 44, p. 9691–9698, https:// doi .org /10 .1002 /2017GL074916 .

Melgar, D., Crowell, B.W., Bock, Y., and Haase, J.S., 2013, Rapid modeling of the 2011 Mw 9.0 Tohoku-oki earthquake with seismogeodesy: Geophysical Research Letters, v. 40, p. 2963–2968, https:// doi .org /10 .1002 /grl .50590 .

Melgar, D., Fan, W.Y., Riquelme, S., Geng, J.H., Liang, C.R., Fuentes, M., Vargas, G., Allen, R.M., Shearer, P.M., and Fielding, E.J., 2016, Slip segmentation and slow rupture to the trench during the 2015, MW 8.3 Illapel, Chile, earthquake: Geophysical Research Letters, v.  43, p. 961–966, https:// doi .org /10 .1002 /2015GL067369 .

Meng, L.S., Huang, H., Burgmann, R., Ampuero, J.P., and Strader, A., 2015a, Dual megathrust slip behaviors of the 2014 Iquique earthquake sequence: Earth and Planetary Science Letters, v. 411, p. 177–187, https:// doi .org /10 .1016 /j .epsl .2014 .11 .041 .

Meng, Q., Heeszel, D.S., Ye, L., Lay, T., Wiens, D.A., Jia, M., and Cummins, P.R., 2015b, The 3 May 2006 (MW 8.0) and 19 March 2009 (MW 7.6) Tonga earthquakes: Intraslab compressional fault-ing below the megathrust: Journal of Geophysical Research–Solid Earth, v. 120, p. 6297–6316, https:// doi .org /10 .1002 /2015JB012242 .

Métois, M., Vigny, C., Socquet, A., Delorme, A., Morvan, S., Ortega, I., and Valderas-Bermejo, C.M., 2014, GPS-derived interseismic coupling on the subduction and seismic hazards in the Atacama region, Chile: Geophysical Journal International, v. 196, p. 644–655, https:// doi .org /10 .1093 /gji /ggt418 .

Miyazaki, S., and Larson, K.M., 2008, Coseismic and early postseismic slip for the 2003 Tokachi- oki earthquake sequence inferred from GPS data: Geophysical Research Letters, v.  35, L04302, https:// doi .org /10 .1029 /2007GL032309 .

Mochizuki, K., Yamada, T., Shinohara, M., Yamanaka, Y., and Kanazawa, T., 2008, Weak inter-plate coupling by seamounts and repeating M ~ 7 earthquakes: Science, v. 321, p. 1194–1197, https:// doi .org /10 .1126 /science .1160250 .

Moore, J.C., and Saffer, D., 2001, Up-dip limit of the seismogenic zone beneath the accretionary prism of southwest Japan: An effect of diagenetic to low-grade metamorphic processes and increasing effective stress: Geology, v. 29, p. 183–186, https:// doi .org /10 .1130 /0091 -7613 (2001)029 <0183: ULOTSZ>2 .0 .CO;2 .

Moreno, M., Rosenau, M., and Oncken, O., 2010, 2010 Maule earthquake slip correlates with pre-seismic locking of Andean subduction zone: Nature, v. 467, p. 198–202, https:// doi .org /10 .1038 /nature09349 .

Moreno, M., Melnick, D., Rosenau, M., et al., 2012, Toward understanding tectonic control on the Mw 8.8 2010 Maule, Chile, earthquake: Earth and Planetary Science Letters, v. 321, p. 152–165, https:// doi .org /10 .1016 /j .epsl .2012 .01 .006 .

Moreno, M., Haberland, C., Oncken, O., Rietbrock, A., Angiboust, S., and Heidbach, O., 2014, Locking of the Chile subduction zone controlled by fluid pressure before the 2010 earth-quake: Nature Geoscience, v. 7, p. 292–296, https:// doi .org /10 .1038 /ngeo2102 .

Nakajima, J., and Hasegawa, A., 2016, Tremor activity inhibited by well-drained conditions above a megathrust: Nature Communications, v. 7, 13863, https:// doi .org /10 .1038 /ncomms13863 .

Nakamura, M., 2017, Distribution of low-frequency earthquakes accompanying the very low fre-quency earthquakes along the Ryukyu Trench: Earth, Planets, and Space, v. 69, p. 49, https:// doi .org /10 .1186 /s40623 -017 -0632 -4 .

Page 31: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1498Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

Naliboff, J.B., Billen, M.I., Gerya, T., and Saunders, J., 2013, Dynamics of outer-rise faulting in oceanic-continental subduction systems: Geochemistry Geophysics Geosystems, v.  14, p. 2310–2327, https:// doi .org /10 .1002 /ggge .20155 .

Newman, A.V., and Okal, E.A., 1998, Teleseismic estimates of radiated seismic energy: The E/Mo discriminant for tsunami earthquakes: Journal of Geophysical Research, v. 103, p. 26,885–26,898, https:// doi .org /10 .1029 /98JB02236 .

Newman, A.V., Hayes, G., Wei, Y., and Convers, J., 2011, The 25 October 2010 Mentawai tsunami earthquake, from real-time discriminants, finite-fault rupture, and tsunami excitation: Geo-physical Research Letters, v. 38, L05302, https:// doi .org /10 .1029 /2010GL046498 .

Nishenko, S.P., 1991, Circum-Pacific seismic potential: 1989–1999: Pure and Applied Geophysics, v. 135, p. 169–259, https:// doi .org /10 .1007 /BF00880240 .

Nishikawa, T., and Ide, S., 2014, Earthquake size distribution in subduction zones linked to slab buoyancy: Nature Geoscience, v. 7, p. 904–908, https:// doi .org /10 .1038 /ngeo2279 .

Nishizawa, A., Kaneda, K., Watanabe, N., and Oikawa, M., 2009, Seismic structure of the sub-ducting seamounts on the trench axis: Erimo Seamount and Daiichi-Kashima Seamount, northern and southern ends of the Japan Trench: Earth, Planets, and Space, v. 61, p. e5–e8, https:// doi .org /10 .1186 /BF03352912 .

Obana, K., and Kodaira, S., 2009, Low-frequency tremors associated with reverse faults in a shallow accretionary prism: Earth and Planetary Science Letters, v. 287, p. 168–174, https:// doi .org /10 .1016 /j .epsl .2009 .08 .005 .

Obara, K., 2002, Nonvolcanic deep tremor associated with subduction in southwest Japan: Sci-ence, v. 296, p. 1679–1681, https:// doi .org /10 .1126 /science .1070378 .

Obara, K., and Ito, Y., 2005, Very low frequency earthquakes excited by the 2004 off the Kii Pen-insula earthquakes: A dynamic deformation process in the large accretionary prism: Earth, Planets, and Space, v. 57, p. 321–326, https:// doi .org /10 .1186 /BF03352570 .

Obara, K., Hirose, H., Yamamizu, F., and Kasahara, K., 2004, Episodic slow slip events accompa-nied by non-volcanic tremors in southwest Japan subduction zone: Geophysical Research Letters, v. 31, L23602, https:// doi .org /10 .1029 /2004GL020848 .

Okal, E.A., and Newman, A.V., 2001, Tsunami earthquakes: The quest for a regional signal: Physics of the Earth and Planetary Interiors, v. 124, p. 45–70, https:// doi .org /10 .1016 /S0031 -9201 (01)00187 -X .

Ozawa, S., Nishimura, T., Suito, H., Kobayashi, T., Tobita, M., and Imakiire, T., 2011, Coseismic and postseismic slip of the 2011 magnitude-9 Tohoku-oki earthquake: Nature, v. 475, p. 373–376, https:// doi .org /10 .1038 /nature10227 .

Ozawa, S., Nishimura, T., Munekane, H., Suito, H., Kobayashi, T., Tobita, M., and Imakiire, T., 2012, Preceding, coseismic, and postseismic slips of the 2011 Tohoku earthquake, Japan: Journal of Geophysical Research, v. 117, B07404, https:// doi .org /10 .1029 /2011JB009120 .

Park, J.-O., Naruse, H., and Bangs, N.L., 2014, Along-strike variations in the Nankai shallow décolle ment properties and their implications for tsunami earthquake generation: Geophysi-cal Research Letters, v. 41, p. 7057–7064, https:// doi .org /10 .1002 /2014GL061096 .

Peacock, S.M., 1990, Fluid processes in subduction zones: Science, v. 248, p. 329–337, https:// doi .org /10 .1126 /science .248 .4953 .329 .

Peacock, S.M., 2009, Thermal and metamorphic environment of subduction-zone episodic tremor and slip: Journal of Geophysical Research, v.  114, B00A07, https:// doi .org /10 .1029 /2008JB005978 .

Peng, Z., and Gomberg, J., 2010, An integrated perspective of the continuum between earthquakes and slow-slip phenomena: Nature Geoscience, v. 3, p. 599–607, https:// doi .org /10 .1038 /ngeo940 .

Perfettini, H., Avouac, J-P., Tavera, H., et al., 2010, Seismic and aseismic slip on the central Peru megathrust: Nature, v. 465, p. 78–81, https:// doi .org /10 .1038 /nature09062 .

Polet, J., and Kanamori, H., 2000, Shallow subduction zone earthquakes and their tsunamigenic potential: Geophysical Journal International, v.  142, p.  684–702, https:// doi .org /10 .1046 /j .1365 -246x .2000 .00205 .x .

Pritchard, M.E., and Fielding, E.J., 2008, A study of the 2006 and 2007 earthquake sequence of Pisco, Peru, with InSAR and teleseismic data: Geophysical Research Letters, v. 35, L09308, https:// doi .org /10 .1029 /2008GL033374 .

Protti, M., McNally, K., Pacheco, J., et al., 1995, The March 25, 1990 (MW = 7.0, ML = 6.8), earth-quake at the entrance of the Nicoya Gulf, Costa Rica: Its prior activity, foreshocks, after-shocks, and triggered seismicity: Journal of Geophysical Research, v. 100, p. 20,345–20,358, https:// doi .org /10 .1029 /94JB03099 .

Protti, M., Gonzalez, V., Newman, A.V., Dixon, T.H., Schwartz, S.Y., Marshall, J.S., Feng, L.J., Walter, J.I., Malservisi, R., and Owen, S.E., 2014, Nicoya earthquake rupture anticipated by

geodetic measurement of the locked plate interface: Nature Geoscience, v.  7, p.  117–121, https:// doi .org /10 .1038 /ngeo2038 .

Raeesi, M., and Atakan, K., 2009, On the deformation cycle of a strongly coupled plate interface: The triple earthquakes of 16 March 1963, 15 November 2006, and 13 January 2007 along the Kurile subduction zone: Journal of Geophysical Research, v. 114, B10301, https:// doi .org /10 .1029 /2008JB006184 .

Ranero, C.R., and von Huene, R., 2000, Subduction erosion along the Middle America conver-gent margin: Nature, v. 404, p. 748–752, https:// doi .org /10 .1038 /35008046 .

Ranero, C.R., Morgan, J.P., McIntosh, K., and Reichert, C., 2003, Bending-related faulting and mantle serpentinization at the Middle America Trench: Nature, v. 425, p. 367–373, https:// doi .org /10 .1038 /nature01961 .

Rhie, J., Dreger, D., Bürgmann, R., and Romanowicz, B., 2007, Slip of the 2004 Sumatra-Anda-man earthquake from joint inversion of long-period global seismic waveforms and GPS static offsets: Bulletin of the Seismological Society of America, v. 97, p. S115–S127, https:// doi .org /10 .1785 /0120050620 .

Rietbrock, A., Ryder, I., Hayes, G., Haberland, C., Comte, D., Roecker, S., and Lyon-Caen, H., 2012, Aftershock seismicity of the 2010 Maule MW = 8.8, Chile, earthquake: Correlation between co-seismic slip models and aftershock distribution?: Geophysical Research Letters, v.  39, L08310, https:// doi .org /10 .1029 /2012GL051308 .

Robinson, D.P., Das, S., and Watts, A.B., 2006, Earthquake rupture stalled by a subducting frac-ture zone: Science, v. 312, p. 1203–1205, https:// doi .org /10 .1126 /science .1125771 .

Rogers, G., and Dragert, H., 2003, Episodic tremor and slip on the Cascadia subduction zone: The chatter of silent slip: Science, v.  300, p.  1942–1943, https:// doi .org /10 .1126 /science .1084783 .

Romano, F., Piatanesi, A., Lorito, S., and Hirata, K., 2010, Slip distribution of the 2003 Tokachi-oki MW 8.1 earthquake from joint inversion of tsunami waveforms and geodetic data: Journal of Geophysical Research–Solid Earth, v. 115, B11313, https:// doi .org /10 .1029 /2009JB006665 .

Romano, F., Trasatti, E., Lorito, S., Piromallo, C., Piatanesi, A., Ito, Y., Zhao, D., Hirata, K., Lanu-cara, P., and Cocco, M., 2014, Structural control on the Tohoku earthquake rupture process investigated by 3D FEM, tsunami and geodetic data: Scientific Reports, v. 4, p. 5631, https:// doi .org /10 .1038 /srep05631 .

Ruff, L.J., 1989, Do trench sediments affect great earthquake occurrence in subduction zones?: Pure and Applied Geophysics, v. 129, p. 263–282, https:// doi .org /10 .1007 /BF00874629 .

Ruiz, S., Metois, M., Fuenzalida, A., Ruiz, J., Leyton, F., Grandin, R., Vigny, C., Madariaga, R., and Campos, J., 2014, Intense foreshocks and a slow slip event preceded the 2014 Iquique Mw 8.1 earthquake: Science, v. 345, p. 1165–1169, https:// doi .org /10 .1126 /science .1256074 .

Rushing, T.M., and Lay, T., 2012, Analysis of seismic magnitude differentials (mb-Mw) across megathrust faults in the vicinity of recent great earthquakes: Earth, Planets, and Space, v. 64, p. 1199–1207, https:// doi .org /10 .5047 /eps .2012 .08 .006 .

Saffer, D.M., and Bekins, B.A., 2002, Hydrologic controls on the morphology and mechanics of accretionary wedges: Geology, v. 30, p. 271–274, https:// doi .org /10 .1130 /0091 -7613 (2002)030 <0271: HCOTMA>2 .0 .CO;2 .

Saffer, D.M., and Tobin, H.J., 2011, Hydrogeology and mechanics of subduction zone forearcs: Fluid flow and pore pressure: Annual Review of Earth and Planetary Sciences, v. 39, p. 157–186, https:// doi .org /10 .1146 /annurev -earth -040610 -133408 .

Saffer, D.M., and Wallace, L.M., 2015, The frictional, hydrologic, metamorphic and thermal habi-tat of shallow slow earthquakes: Nature Geoscience, v. 8, p. 594–600, https:// doi .org /10 .1038 /ngeo2490 .

Satake, K., Fujii, Y., Harada, T., and Namegaya, Y., 2013, Time and space distribution of coseismic slip of the 2011 Tohoku earthquake as inferred from tsunami waveform data: Bulletin of the Seismological Society of America, v. 103, no. 2B, p. 1473–1492, https:// doi .org /10 .1785 /0120120122 .

Sato, M., Ishikawa, T., Ujihara, N., Yoshida, S., Fujita, M., Mochizuki, M., and Asada, A., 2011, Dis-placement above the hypocenter of the 2011 Tohoku-oki earthquake: Science, v. 332, p. 1395, https:// doi .org /10 .1126 /science .1207401 .

Scholl, D.W., Kirby, S.H., van Huene, R., Ryan, H., Wells, R.E., and Geist, E.L., 2015, Great mega-thrust earthquake and subduction of excess and bathymetrically smooth seafloor: Geo-sphere, v. 11, p. 236–265, https:// doi .org /10 .1130 /GES01079 .1 .

Scholz, C.H., and Campos, J., 1995, On the mechanism of seismic decoupling and back-arc spreading in subduction zones: Journal of Geophysical Research, v. 100, p. 22,103–22,115, https:// doi .org /10 .1029 /95JB01869 .

Page 32: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1499Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

Scholz, C.H., and Campos, J., 2012, The seismic coupling of subduction zones revisited: Journal of Geophysical Research, v. 117, B05310, https:// doi .org /10 .1029 /2011JB009003 .

Scholz, C.H., and Small, C., 1997, The effect of seamount subduction on seismic coupling: Geol-ogy, v. 25, p. 487–490, https:// doi .org /10 .1130 /0091 -7613 (1997)025 <0487: TEOSSO>2 .3 .CO;2 .

Schorlemmer, D., and Wiemer, S., 2005, Microseismicity data forecast rupture area: Nature, v. 434, p. 1086, https:// doi .org /10 .1038 /4341086a .

Schorlemmer, D., Wiemer, S., and Wyss, M., 2005, Variation in earthquake size distribution across different stress regimes: Nature, v. 437, p. 539–542, https:// doi .org /10 .1038 /nature04094 .

Schwartz, S.Y., and Rokosky, J.M., 2007, Slow slip events and seismic tremor at circum- Pacific subduction zones: Reviews of Geophysics, v.  45, RG3004, https:// doi .org /10 .1029 /2006RG000208 .

Seno, T., 2017, Subducted sediment thickness and MW 9 earthquakes: Journal of Geophysical Research–Solid Earth, v. 122, p. 470–491, https:// doi .org /10 .1002 /2016JB013048 .

Shearer, P.M., and Bürgmann, R., 2010, Lessons learned from the 2004 Sumatra-Andaman megathrust rupture: Annual Review of Earth and Planetary Sciences, v.  38, p.  103–131, https:// doi .org /10 .1146 /annurev -earth -040809 -152537 .

Shelly, D.R., Beroza, G.C., Ide, S., and Nakamula, S., 2006, Low-frequency earthquakes in Shi-koku, Japan, and their relationship to episodic tremor and slip: Nature, v. 442, p. 188–191, https:// doi .org /10 .1038 /nature04931 .

Shelly, D.R., Beroza, G.C., and Ide, S., 2007, Non-volcanic tremor and low-frequency earthquake swarms: Nature, v. 446, p. 305–307, https:// doi .org /10 .1038 /nature05666 .

Sieh, K., Natawidjaja, D.H., Meltzner, A.J., et  al., 2008, Earthquake supercycles inferred from sea-level changes recorded in the corals of west Sumatra: Science, v. 322, no. 5908, p. 1674–1678, https:// doi .org /10 .1126 /science .1163589 .

Simons, M., Minson, S.E., Sladen, A., et al., 2011, The 2011 magnitude 9.0 Tohoku-oki earthquake: Mosaicking the megathrust from seconds to centuries: Science, v. 332, p. 1421–1425, https:// doi .org /10 .1126 /science .1206731 .

Singh, S.K., and Suárez, G., 1988, Regional variation in the number of aftershocks (mb ≥ 5) of large, subduction-zone earthquakes (Mw ≥ 7): Bulletin of the Seismological Society of Amer-ica, v. 78, p. 230–242.

Sladen, A., and Trevisan, J., 2018, Shallow megathrust earthquake ruptures betrayed by their outer-trench aftershocks signature: Earth and Planetary Science Letters, v. 483, p. 105–113, https:// doi .org /10 .1016 /j .epsl .2017 .12 .006 .

Sladen, A., Tavera, H., Simons, M., Avouac, J.-P., Konca, A.O., Perfettini, H., Audin, L., Fielding, E.J., Ortega, F., and Cavagnoud, R., 2010, Source model of the 2007 MW 8.0 Pisco, Peru, earthquake—Implications for seismogenic behavior of subduction megathrusts: Journal of Geophysical Research, v. 115, B02405, https:// doi .org /10 .1029 /2009JB006429 .

Socquet, A., Valdes, J.P., Jara, J., Cotton, F., Walpersdorf, A., Cotte, N., Specht, S., Ortega-Cula-ciati, F., Carrizo, D., and Norabuena, E., 2017, An 8 month slow slip event triggers progressive nucleation of the 2014 Chile megathrust: Geophysical Research Letters, v. 44, p. 4046–4053, https:// doi .org /10 .1002 /2017GL073023 .

Song, T.R.A., and Simons, M., 2003, Large trench-parallel gravity variations predict seismogenic behavior in subduction zones: Science, v.  301, p.  630–633, https:// doi .org /10 .1126 /science .1085557 .

Song, T.R.A., Helmberger, D.V., Brudzinski, M.R., Clayton, R.W., Davis, P., Perez-Campos, X., and Singh, S.K., 2009, Subducting slab ultra-slow velocity layer coincident with silent earth-quakes in southern Mexico: Science, v.  324, p.  502–506, https:// doi .org /10 .1126 /science .1167595 .

Sparkes, R., Tilmann, F., Hovius, N., and Hillier, J., 2010, Subducted seafloor relief stops rupture in South American great earthquakes: Implications for rupture behaviour in the 2010 Maule, Chile, earthquake: Earth and Planetary Science Letters, v. 298, p. 89–94, https:// doi .org /10 .1016 /j .epsl .2010 .07 .029 .

Spence, W., Mendoza, C., Engdahl, E.R., Choy, G.L., and Norabuena, E., 1999, Seismic subduc-tion of the Nazca Ridge as shown by the 1996–97 Peru earthquakes: Pure and Applied Geo-physics, v. 154, p. 753–776, https:// doi .org /10 .1007 /s000240050251 .

Stankova-Pursley, J., Bilek, S.L., Phillips, W.S., and Newman, A.V., 2011, Along-strike variations of earthquake apparent stress at the Nicoya Peninsula, Costa Rica, subduction zone: Geo-chemistry Geophysics Geosystems, v. 12, Q08002, https:// doi .org /10 .1029 /2011GC003558 .

Suárez, G., and Albini, P., 2009, Evidence for great tsunamigenic earthquakes (M 8.6) along the Mexican subduction zone: Bulletin of the Seismological Society of America, v. 99, p. 892–896, https:// doi .org /10 .1785 /0120080201 .

Taylor, F.W., Briggs, R.W., Frohlich, C., Brown, A., Hornbach, M., Papabatu, A.K., Meltzner, A.J., and Billy, D., 2008, Rupture across arc segment and plate boundaries in the 1 April 2007 Solomons earthquake: Nature Geoscience, v. 1, p. 253–257, https:// doi .org /10 .1038 /ngeo159 .

Tilmann, F., Zhang, Y., Moreno, M. et al., 2016, The 2015 Illapel earthquake, central Chile: A type case for a characteristic earthquake?: Geophysical Research Letters, v. 43, p. 574–583, https:// doi .org /10 .1002 /2015GL066963 .

Tobin, H.J., and Saffer, D.M., 2009, Elevated fluid pressure and extreme mechanical weakness of a plate boundary thrust, Nankai Trough subduction zone: Geology, v. 37, p. 679–682, https:// doi .org /10 .1130 /G25752A .1 .

Tormann, T., Enescu, B., Woessner, J., and Wiemer, S., 2015, Randomness of megathrust earth-quakes implied by rapid stress recovery after the Japan earthquake: Nature Geoscience, v. 8, p. 152–158, https:// doi .org /10 .1038 /ngeo2343 .

Tréhu, A.M., Blakely, R.J., and Williams, M.C., 2012, Subducted seamounts and recent earth-quakes beneath the central Cascadia forearc: Geology, v. 40, p. 103–106, https:// doi .org /10 .1130 /G32460 .1 .

Uchida, N., Shimamura, K., Matsuzawa, T., and Okada, T., 2015, Postseismic response of repeat-ing earthquakes around the 2011 Tohoku-oki earthquake: Moment increases due to the fast loading rate: Journal of Geophysical Research, v.  120, p.  259–274, https:// doi .org /10 .1002 /2013JB010933 .

Uchide, T., Shearer, P.M., and Imanishi, K., 2014, Stress drop variations among small earthquakes before the 2011 Tohoku-oki, Japan, earthquake and implications for the main shock: Journal of Geophysical Research, v. 119, p. 7164–7174, https:// doi .org /10 .1002 /2014JB010943 .

Utsu, T., 1965, A method for determining the value of b in a formula log n = a – bm showing the magnitude-frequency relation for earthquakes: Geophysical Bulletin of the Hokkaido Uni-versity, v. 13, p. 99–103.

Uyeda, S., and Kanamori, H., 1979, Back-arc opening and mode of subduction: Journal of Geo-physical Research, v. 84, no. B3, p. 1049–1061, https:// doi .org /10 .1029 /JB084iB03p01049 .

Vallée, M., and Douet, V., 2016, A new database of source time functions (STFs) extracted from the SCARDEC method: Physics of the Earth and Planetary Interiors, v. 257, p. 149–157, https:// doi .org /10 .1016 /j .pepi .2016 .05 .012 .

Venkataraman, A., and Kanamori, H., 2004, Observational constraints on the fracture energy of subduction zone earthquakes: Journal of Geophysical Research, v. 109, B05302, https:// doi .org /10 .1029 /2003JB002549 .

Vigny, C., Socquet, A., Peyrat, S., et al., 2011, The 2010 MW 8.8 Maule megathrust earthquake of central Chile, monitored by GPS: Science, v.  332, p.  1417–1421, https:// doi .org /10 .1126 /science .1204132 .

Wallace, L.M., Beavan, J., McCaffrey, R., and Darby, D., 2004, Subduction zone coupling and tectonic block rotations in the North Island, New Zealand: Journal of Geophysical Research, v. 109, B12406, https:// doi .org /10 .1029 /2004JB003241 .

Wallace, L.M., Webb, S.C., Ito, Y., Mochizuki, K., Hino, R., Henrys, S., Schwartz, S.Y., and Sheehan, A.F., 2016, Slow slip near the trench at the Hikurangi subduction zone, New Zealand: Science, v. 352, p. 701–704, https:// doi .org /10 .1126 /science .aaf2349 .

Wang, K., and Bilek, S.L., 2011, Do subducting seamounts generate or stop large earthquakes?: Geology, v. 39, p. 819–822, https:// doi .org /10 .1130 /G31856 .1 .

Wang, K., and Bilek, S.L., 2014, Invited review paper: Fault creep caused by subduction of rough seafloor relief: Tectonophysics, v. 610, p. 1–24, https:// doi .org /10 .1016 /j .tecto .2013 .11 .024 .

Wang, K., and Dixon, T., 2004, “Coupling” semantics and science in earthquake research: Eos (Transactions, American Geophysical Union), v. 85, p. 180, https:// doi .org /10 .1029 /2004EO180005 .

Wells, R.E., Blakely, R.J., Sugiyama, Y., Scholl, D.W., and Dinterman, P.A., 2003, Basin-centered asperities in great subduction zone earthquakes: A link between slip, subsidence, and sub-duction erosion?: Journal of Geophysical Research, v. 108, p. 2507, https:// doi .org /10 .1029 /2002JB002072 .

Wells, R.E., Blakely, R.J., Wech, A.G., McCrory, P.A., and Michael, A., 2017, Cascadia subduction tremor muted by crustal faults: Geology, v. 45, p. 515–518, https:// doi .org /10 .1130 /G38835 .1 .

Wetzler, N., Brodsky, E., and Lay, T., 2016, Regional and stress drop effects on aftershock pro-ductivity of large megathrust earthquakes: Geophysical Research Letters, v. 43, p. 12,012–12,020, https:// doi .org /10 .1002 /2016GL071104 .

Wetzler, N., Lay, T., Brodsky, E.E., and Kanamori, H., 2017, Rupture-depth-varying seismicity pat-terns for major and great (MW ≥ 7.0) megathrust earthquakes: Geophysical Research Letters, v. 44, p. 9663–9671, https:// doi .org /10 .1002 /2017GL074573 .

Page 33: GEOSPHERE Subduction zone megathrust earthquakesthorne/TL.pdfs/BL... · 2019-07-24 · Research Paer GEOSPHERE | Volume 14 Number 4 Bi nd | Subduction on mtut tu 1468 Subduction zone

Research Paper

1500Bilek and Lay | Subduction zone megathrust earthquakesGEOSPHERE | Volume 14 | Number 4

Wetzler, N., Lay, T., Brodsky, E.E., and Kanamori, H., 2018, Systematic deficiency of aftershocks in areas of high coseismic slip for large subduction zone earthquakes: Scientific Advances, v. 4, p. eaao3225, https:// doi .org /10 .1126 /sciadv .aao3225.

Yamamoto, Y., Obana, K., Kodaira, S., Hino, R., and Shinohara, M., 2014, Structural heterogene-ities around the megathrust zone of the 2011 Tohoku earthquake from tomographic inver-sion of onshore and offshore seismic observations: Journal of Geophysical Research–Solid Earth, v. 119, p. 1165–1180, https:// doi .org /10 .1002 /2013JB010582 .

Yamashita, Y., Yakiwara, H., Asano, Y., et  al., 2015, Migrating tremor off southern Kyushu as evidence for slow slip of a shallow subduction interface: Science, v. 348, p. 676–679, https:// doi .org /10 .1126 /science .aaa4242 .

Yamazaki, Y., Cheung, K.F., and Lay, T., 2018, A self-consistent fault slip source model for the 2011 Tohoku earthquake and tsunami: Journal of Geophysical Research–Solid Earth, v. 123, p. 1435–1458, https:// doi .org /10 .1002 /2017JB014749 .

Yang, H., Liu, Y., and Lin, J., 2013, Geometrical effects of a subducted seamount on stopping megathrust ruptures: Geophysical Research Letters, v.  40, p.  2011–2016, https:// doi .org /10 .1002 /grl .50509 .

Ye, L., Kanamori, H., Avouac, J.P., Li, L.Y., Cheung, K.F., and Lay, T., 2016a, The 16 April 2016, MW 7.8 (MS 7.5) Ecuador earthquake: A quasi-repeat of the 1942 MS 7.5 earthquake and partial re-rupture of the 1906 MS 8.6 Colombia-Ecuador earthquake: Earth and Planetary Science Letters, v. 454, p. 248–258, https:// doi .org /10 .1016 /j .epsl .2016 .09 .006 .

Ye, L., Lay, T., Kanamori, H., and Rivera, L., 2016b, Rupture characteristics of major and great (MW ≥ 7.0) megathrust earthquakes from 1990 to 2015: 1. Source parameter scaling re-lationships: Journal of Geophysical Research, v.  121, p.  826–844, https:// doi .org /10 .1002 /2015JB012426 .

Ye, L., Lay, T., Kanamori, H., and Rivera, L., 2016c, Rupture characteristics of major and great (MW ≥ 7.0) megathrust earthquakes from 1990 to 2015: 2. Depth dependence: Journal of Geo physi-cal Research, v. 121, p. 845–863, https:// doi .org /10 .1002 /2015JB012427 .

Yin, J.X., Yang, H.F., Yao, H.J., and Weng, H.H., 2016, Coseismic radiation and stress drop during the 2015 Mw 8.3 Illapel, Chile, megathrust earthquake: Geophysical Research Letters, v. 43, p. 1520–1528, https:// doi .org /10 .1002 /2015GL067381 .

Yokota, Y., Ishikawa, T., Watanabe, S., Tashiro, T., and Asada, A., 2016, Seafloor geodetic con-straints on interplate coupling of the Nankai Trough megathrust zone: Nature, v. 534, p. 374–377, https:// doi .org /10 .1038 /nature17632 .

Yoshimoto, M., Watada, S., Fujii, Y., and Satake, K., 2016, Source estimate and tsunami fore-cast from far-field deep-ocean tsunami waveforms—The 27 February 2010 MW 8.8 Maule earthquake: Geophysical Research Letters, v.  43, p.  659–665, https:// doi .org /10 .1002 /2015GL067181 .

Yue, H., Lay, T., Schwartz, S.Y., Rivera, L., Protti, M., Dixon, T.H., Owen, S., and Newman, A.V., 2013, The 5 September 2012 Nicoya, Costa Rica, MW 7.6 earthquake rupture process from joint inversion of high-rate GPS, strong-motion, and teleseismic P-wave data and its rela-tionship to adjacent plate boundary interface properties: Journal of Geophysical Research, v. 118, p. 5453–5466, https:// doi .org /10 .1002 /jgrb .50379 .

Yue, H., Lay, T., Rivera, L., Bai, Y., Yamazaki, Y., Cheung, K.F., Hill, E.M., Sieh, K., Kongko, W., and Muhari, A., 2014a, Rupture process of the 2010 Mw 7.8 Mentawai tsunami earthquake from joint inversion of near-field hr-GPS and teleseismic body wave recordings constrained by tsunami observations: Journal of Geophysical Research, v. 119, p. 5574–5593, https:// doi .org /10 .1002 /2014JB011082 .

Yue, H., Lay, T., Rivera, L., An, C., Vigny, C., Tong, X., and Báez Soto, J.C., 2014b, Localized fault slip to the trench in the 2010 Maule, Chile, MW  = 8.8 earthquake from joint inversion of high-rate GPS, teleseismic body waves, InSAR, campaign GPS, and tsunami observations: Journal of Geophysical Research, v. 119, p. 7786–7804, https:// doi .org /10 .1002 /2014JB011340 .

Zhao, D., Huang, Z., Umino, N., Hasegawa, A., and Kanamori, H., 2011, Structural heterogeneity in the megathrust zone and mechanism of the 2011 Tohoku-oki earthquake (MW 9.0): Geo-physical Research Letters, v. 38, L17308, https:// doi .org /10 .1029 /2011GL048408 .