You are on page 1of 11

Earth-Science Reviews 141 (2015) 4555

Contents lists available at ScienceDirect

Earth-Science Reviews
journal homepage: www.elsevier.com/locate/earscirev

Coseismic slip on shallow dcollement megathrusts: implications for


seismic and tsunami hazard
Judith Hubbard , Sylvain Barbot, Emma M. Hill, Paul Tapponnier
Earth Observatory of Singapore, 50 Nanyang Avenue, Nanyang Technological University, 639798, Singapore

a r t i c l e

i n f o

Article history:
Received 24 July 2014
Accepted 8 November 2014
Available online 15 November 2014
Keywords:
Earthquakes
Dcollement
Subduction zone
Fold-and-thrust belt
Seismogenic zone
Faults

a b s t r a c t
For years, many studies of subduction zones and on-land fold-and-thrust belts have assumed that the frontal portions of accretionary prisms are too weak to rupture coseismically and must therefore be fully creeping. We present a series of examples, both on-land and offshore, demonstrating that in many cases, shallow dcollements are
capable of large, coseismic slip events that rupture to the toes of the fault systems. Some of these events are associated with ruptures that initiate down-dip, while others appear to be limited to the frontal, shallow portion of
the wedge.
We suggest that this behavior is not limited to the examples described here, but rather is common to many (perhaps most) accretionary wedges and fold-and-thrust belts around the world. Indeed, there may be many other
examples of similar earthquakes, where existing data cannot constrain slip at the toe. We do not characterize
the regions and events described here as unusual, as they encompass a wide range of settings. This study indicates that there is an urgent need to reevaluate seismic and tsunami hazard in fold-and-thrust belts and subduction zones around the world, allowing for the possibility of shallow dcollement rupture.
2014 Elsevier B.V. All rights reserved.

Contents
1.
2.

3.

4.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Evidence for shallow slip in megathrust earthquakes . . . . . . . . . . . . . . . . . . .
2.1.
The Himalaya
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.
Bolivia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.
Western Taiwan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.
Santa Barbara Channel, Southern California . . . . . . . . . . . . . . . . . . . .
2.5.
Sumatra subduction zone, Mentawai segment . . . . . . . . . . . . . . . . . . .
2.6.
Sumatra subduction zone, northern segment . . . . . . . . . . . . . . . . . . .
2.7.
Java subduction zone . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.8.
Japan Trench . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.9.
Kuril Trench . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.10.
Japan, Nankai Trough
. . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.11.
Solomon Islands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.12.
Middle America megathrust . . . . . . . . . . . . . . . . . . . . . . . . . .
2.13.
Alaskan Trench . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.14.
Peru . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.15.
Chile . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.16.
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
How can weak decollements behave seismically? . . . . . . . . . . . . . . . . . . . . .
3.1.
The state of stress within the wedge, and preferred slip planes at geological time scales
3.2.
Variations in strength through the earthquake cycle . . . . . . . . . . . . . . . .
Consequences of seismogenic decollements . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

Corresponding author. Tel.: +65 9855 2730, +65 6592 7537.


E-mail addresses: judith.a.hubbard@gmail.com (J. Hubbard), sylbar.vainbot@gmail.com (S. Barbot), ehill@ntu.edu.sg (E.M. Hill), tappon@ntu.edu.sg (P. Tapponnier).

http://dx.doi.org/10.1016/j.earscirev.2014.11.003
0012-8252/ 2014 Elsevier B.V. All rights reserved.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

46
47
47
48
49
49
50
50
50
50
50
50
51
51
51
51
51
51
51
52
53
53

46

J. Hubbard et al. / Earth-Science Reviews 141 (2015) 4555

5.
Discussion and conclusions
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
References
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

1. Introduction
Many regions of plate convergence are underlain by dcollement
megathrusts. They form the base of both accretionary wedges and foldand-thrust belts. These faults may extend laterally for hundreds or thousands of kilometers, and downdip for tens to hundreds of kilometers. Traditionally, estimates of seismic hazard have assumed that these faults slip
aseismically, without radiating signicant seismic energy (e.g., Pacheco
et al., 1993; Hyndman et al., 1997; Oleskevich et al., 1999). However, in
several recent cases, shallow dcollements have been shown to slip in
large, discrete events (e.g., 2011 Mw 9.0 Tohoku-Oki earthquake, Japan;
2010 Mw 7.8 Mentawai earthquake, Indonesia; 1999 Mw 7.6 Chi-Chi
earthquake, Taiwan). On land, dcollements frequently lie along the borders of large, populated basins, and therefore pose an important seismic
hazard, threatening such large cities as Dhaka (Bangladesh), Chengdu
(China), Baghdad (Iraq), and Delhi (India). Offshore, they form the
lower boundaries of accretionary prisms in subduction zones, and should
be considered in seismic and tsunami hazard assessment in subduction
zones around the world.
Unlike reverse faults, which form at dips of ~ 2060 according to
both observation and theory, dcollements dip gently, at angles of b1
10 (Davis et al., 1983). This is possible because these faults take advantage of preexisting weaknesses in the rock, forming along stratigraphic
horizons with weak materials like salt or shale (Suppe, 2007; Hubbard
et al., 2010), in some cases with high pore pressures (Behrmann et al.,
1988; Bilotti and Shaw, 2005; Cubas et al., 2013). The existence and
long-term deformation associated with dcollements is understood not
only through observation (e.g., Ye et al., 1997; Adam et al., 2004;
Moore et al., 2009; Morley et al., 2011), but also through laboratory
(e.g., Malaveille, 2010; Graveleau et al., 2012), computer (e.g., Strayer
et al., 2001; Burbidge and Braun, 2002), and theoretical modeling
(Davis et al., 1983; Dahlen et al., 1984; Dahlen, 1990). Large, active
dcollements are known to exist in regions both onshore (Himalayas;
Taiwan; Bolivia; Bangladesh; Sichuan, China;) and offshore (Sumatra,
Java, Japan, Peru, Cascadia, Antilles, Makran, Guatemala) (Davis et al.,
1983; Bilotti and Shaw, 2005; Hubbard et al., 2010; Morley et al., 2011).
Many studies of seismic hazard have assumed that because
dcollements are weak, they are unable to support large stresses and
store sufcient elastic energy to produce hazardous earthquakes, but
must rather be fully creeping, generating only small and microearthquakes (e.g., Byrne et al., 1988; Hyndman et al., 1997). Studies of
subduction zones have generally observed an updip limit to interplate
seismicity that persists for several decades (the seismic front, Byrne
et al., 1988; Fig. 1), which has led to the application of the term aseismic
to the portion of the dcollement underlying the accretionary prism. Although this term is correctly applied in that we observe little seismicity
in this region, it has also been taken to mean that this part of the wedge
never slips in association with moderate to large earthquakes an assumption rather than an observation.
Experimental studies of clay and gouge materials show that bare
rock surfaces and thin gouge layers exhibit potentially unstable
velocity-weakening behavior, while slip within thick gouge shows
velocity-strengthening behavior (Marone and Scholz, 1988). The
updip limit has been inferred to be associated with a zone of thick,
velocity-strengthening material in the accretionary prism that resists
rapid rupture (Marone and Scholz, 1988). Thus, the absence of observed
moderate to large earthquakes in this region has been used to infer a slip
behavior that in turn has been used to justify a stratigraphic model that
is consistent with creeping behavior.

As a consequence, until recently, many models of both coseismic slip


and interseismic coupling on subduction zones started with the assumption that there was no coseismic slip at the tip of the wedge, and
that the region updip of the seismic front was fully creeping (Byrne
et al., 1988; Hyndman et al., 1997; Chlieh et al., 2007, 2008; Loveless
and Meade, 2009). Inversions for coseismic slip or interseismic coupling
often appear to support the idea that the shallow region is creeping.
However, these inversions often force the near-trench area of the fault
to creep during the interseismic period or prevent it from slipping during earthquakes. Inferences about the kinematic behavior of the toe of
the prism are therefore not usually directly supported by data, as
discussed by Rhie et al. (2007) and Loveless and Meade (2011).
We agree that dcollements appear to have long-term weak behavior (Suppe, 2007), and that there is often a strong drop-off in recorded
seismicity updip of the seismic front. However, neither of these observations is a compelling reason to infer creeping behavior. Seismic behavior
is possible for a weak fault: for example with a low effective conning
pressure of 50 MPa and a low effective coefcient of friction of 0.1,
there is enough frictional resistance for a complete coseismic stress
drop of 5 MPa, a value larger than that for typical inter-plate earthquakes (Venkataraman and Kanamori, 2004). The opposite can also be
true: a creeping fault may not be weak, but rather be creeping at a
higher shear stress than a stick-slip fault, averaged over the seismic
cycle. And indeed, contrary to the suggestion that low seismicity indicates creeping behavior, current understanding of fault zones rather
suggests the opposite: we often expect to see seismicity where faults
are creeping, and little seismicity where they are locked (e.g., Rubin
et al., 1999; Barbot et al., 2013). Further, these observations of limited
seismicity may be dependent on the earthquake cycle: in Sumatra, the
seismic front is clear prior to the year 2000, but following the Mw 7.9
earthquake in Southern Sumatra in 2000 (and the great earthquakes
in 2004 and 2005), many earthquakes were recorded in the frontal
part of the system (Fig. 2: Aceh, Nias/Simeulue, and Mentawai 2010 segments). In addition to the temporal change in seismicity in Sumatra, we
also observe tremendous spatial variability, with portions of the
megathrust exhibiting minimal seismicity everywhere, other showing
a clear seismic front, and yet others generating earthquakes everywhere
up to 300 km from the trench (Fig. 2). Thus, a single dcollement may
have extremely variable seismic patterns both spatially and temporally.
As described below, we can nd evidence in many dcollement systems
for large, episodic slip events at the tips of wedge systems, often associated with recorded earthquakes. This indicates that current models of
fully creeping behavior on these systems are awed (Fig. 1).
Several recent earthquakes, including the 2010 Mw 7.8 Mentawai
tsunami earthquake, Indonesia, and the 2011 Mw 9.0 Tohoku-Oki earthquake, Japan have prompted a reevaluation of the assumption of creeping behavior near the trench (Lay and Bilek, 2007; McCaffrey, 2008;
Avouac, 2011; Faulkner et al., 2011; Loveless and Meade, 2011; Hill
et al., 2012; Kozdon and Dunham, 2013) and the development of new
models where creep and coseismic slip occur at different stages of the
earthquake cycle (Noda and Lapusta, 2013). Here, we present a review
of additional data from these and other dcollement systems around
the world, both on- and offshore, that support the inference that they
can rupture in earthquakes associated with large slip at their tips.
Although many studies implicitly treat subduction zones and continental fold-and-thrust belts as different classes, we show throughout
this paper that they are structurally similar (Fig. 3) and exhibit much
the same behavior. This paper brings together perspectives from structural geology, geodesy, and earthquake dynamics, both on-land and

J. Hubbard et al. / Earth-Science Reviews 141 (2015) 4555

47

Fig. 1. Schematic model of a subduction zone, showing the locations of different types of earthquakes. (A) Redrawn after Byrne et al. (1988). (B) Our model. We suggest that the
dcollement at the base of the accretionary prism is seismogenic and/or capable of participating in large thrust events on the subduction interface. We suggest that the faults rising
from the shallow dcollement are also capable of participating in earthquakes, either in addition to or instead of rupture all the way to the toe.

offshore. We suggest that exposed fold-and-thrust belts can provide important constraints and observations that can be used to understand the
behavior of subduction zones, where observations are much more
limited.

2. Evidence for shallow slip in megathrust earthquakes


Below, we describe a set of regions with large dcollements showing
evidence for earthquakes that rupture to their tips in large slip events
(Fig. 3; Shaw and Suppe, 1994; Ye et al., 1997; Lav and Avouac, 2000;
Ranero et al., 2000; Yue et al., 2005; Singh et al., 2008; Moore et al.,
2009; Uba et al., 2009; Shulgin et al., 2011; Singh et al., 2011; Hubbard
et al, 2014). This evidence falls primarily into four categories: (1) locking
occurs along the dcollement, with geodetic observations of strain accumulating downdip; (2) slip propagating to the toe of the system over
timescales of hundreds of years to tens of thousands of years, implying
the toe is active; (3) slip that occurs episodically (i.e. not as interseismic
creep), and (4) slip associated with earthquakes, as opposed to just
slow slip events (e.g., Meade and Loveless, 2009).

2.1. The Himalaya


One of the most compelling cases for large earthquakes on
dcollements is in the Himalaya. This ongoing continent-continent
collision is now primarily accommodated along a 2000-km-wide
subhorizontal thrust that dips gently beneath the lesser Himalaya.
In the Nepal Himalaya, the dcollement extends about 90 km
downdip, and then steepens beneath the high Himalaya on a blind
thrust ramp (Schelling and Arita, 1991; Jackson and Bilham, 1994;
Pandey et al., 1995). Much like subduction zones, the Himalaya exhibits a seismic front, with intense microseismicity and frequent
small earthquakes along the ramp, and limited seismicity along the
dcollement itself (Pandey et al., 1995).
Geodetic measurements in the Himalaya demonstrate that the
dcollement is currently locked from the surface to about 100 km
downdip (Bilham et al., 1997; Larson et al., 1999; Jouanne et al., 2004;
Betinelli et al., 2006; Ader et al., 2012). Over the ten thousand year timescale, the intraplate motion is accommodated by migrating slip onto the
presently locked zone, and most of this slip propagates to the toe, based
on uplifted terraces on the frontal fault of the system (Lav and Avouac,

48

J. Hubbard et al. / Earth-Science Reviews 141 (2015) 4555

40

N
60 km
40 km

20 km

a
0

20

pre-2000

a) Aceh segment

200

40 km

92

post-2000

10

km

30

40

b) Nias/Simeulue segment

30
20
10

96

Inv
e
ac stig
a
tur
e Z tor
on
e

Fr

c) Mentawai seismic gap

30

40

Number of earthquakes

0
Meg
athr us t

94

S u n da

20
10
0
40

d) Pagai

30
20
10

98

Sumatra

40

e) Mentawai 2010 patch

30
20

f
8

10
0

10

40

f) Enggano segment

30
6

20

10

10

10
4

10

Date of earthquakes

0
0
2002

2004

2006

2008

2010

50

100

150

200

250

300

Distance from the trench (km)

2012

Fig. 2. (Left) Map of Sumatra showing seismicity, ltered for thrust events (data extracted from the Global CMT catalog). Contours for the subduction zone are shown as thin black lines. Dashed
lines show boundaries of regions for seismicity histograms (a-f; shown on the right). Focal mechanisms for earthquakes M N = 7.0 are shown. (Right) Histograms of seismicity along the Sumatran megathrust seismicity (red) prior to the 2000 Mw 7.9 earthquake in Southern Sumatra and the later great earthquakes, and (black) from 20002014, for six different portions of the
megathrust. Seismicity is binned into 10-km stripes down-dip. The megathrust demonstrates strong variability in seismicity, both temporally and spatially (along strike and down dip).

2000). This happens in discrete events with large amounts of slip at the
toe that can be tied to historical large earthquakes (Sapkota et al., 2013;
Bollinger et al., 2014). Thus, in Nepal it appears not only that the
dcollement is slipping in large events that reach the surface, but also
that these events are radiating signicant seismic energy.

2.2. Bolivia
The eastern side of the Central Andes forms a backarc fold-andthrust belt underlain by a shallowly dipping dcollement. Shortening
across the region is estimated at ~ 7-13 mm/yr from both geological

J. Hubbard et al. / Earth-Science Reviews 141 (2015) 4555

49

Fig. 3. Schematic diagrams of continental fold-and-thrust belts and subduction zones from around the world that show evidence of shallow dcollement earthquakes, shown at a xed
scale. Most of the regions discussed in the text are represented here.

and geodetic data (Uba et al., 2009; Brooks et al., 2011b). Like in the
Himalaya, geodetic measurements suggest that the dcollement is locked
from the toe to about 100 km downdip (Brooks et al., 2011b). In addition,
Brooks et al. (2011a) have found geological evidence for a surfacerupturing event with at least 7 m of slip at the range-front fault at the
tip of the system.

2.3. Western Taiwan


The 1999 Mw 7.6 Chi-Chi earthquake represents a clear example in
which a shallow, bedding-parallel dcollement ruptured with large surface slip (3-10 m) (Lee and Ma, 1999; Johnson et al., 2001; Yue et al.,
2005). Although the rupture involved a complex geometry and fault
properties (Ma et al., 2000, 2003), it is clear that the earthquake produced
signicant coseismic slip (and afterslip) on the bedding-parallel
Chelungpu-Sanyi thrust system, which extends as a shallow dcollement

(5-8 km deep) beneath western Taiwan for tens of kilometers (Yue et al.,
2005; Rousset et al., 2012).
2.4. Santa Barbara Channel, Southern California
The southwestern margin of the Transverse Ranges extends offshore
into the Santa Barbara Channel along a set of bedding-parallel
dcollements at 36 km depth, imaged by seismic reection data
(Shaw and Suppe, 1994). Syntectonic sediments deposited on these
structures produce distinctive growth triangles that record active slip
in the Quaternary accommodating ~ 45% of the geodetically measured
shortening rate across the channel. North of the toe of the dcollement,
the Ventura-Pitas Point fault splays upward, producing the Ventura Avenue anticline (Hubbard et al., 2014); recent terrace measurements
demonstrate that this fault produces large, episodic uplift events
(Rockwell, 2011). These large uplift events would require rupture of
not only the Ventura-Pitas Point fault, but also the dcollement and

50

J. Hubbard et al. / Earth-Science Reviews 141 (2015) 4555

other fault systems along strike (Hubbard et al., 2014). We therefore


conclude that shortening is at least in part accommodated by large, episodic events on the dcollement system.
2.5. Sumatra subduction zone, Mentawai segment
The 2010 Mw 7.8 Mentawai earthquake produced a tsunami that
was much larger than expected based on the seismic magnitude, leading to its classication as a tsunami earthquake (Kanamori, 1972).
Hill et al. (2012) used GPS data and a tsunami eld survey to demonstrate that the earthquake must have been associated with high fault
slip at shallow depths close to the oceanic trench, with most of the
slip at depths shallower than 6 km. Yue et al. (2014) used a joint inversion of high-rate GPS and teleseismic data to conrm this result, placing
even higher slip (up to 23 m) in the shallowest section (above 5 km).
Such high slip at the tip of an accretionary prism must be the result of
slip along the dcollement at its base. This event provides a possible example of a large earthquake that initiated updip of the seismic front,
indicating that the dcollement here may be not only seismic, but
seismogenic. Alternatively, it is possible that the rupture initiated
deeper, but that the deeper slip was low in magnitude compared to
the shallower slip, and is therefore not well constrained. Coral
microatolls in this region also record a slightly deeper, but still shallow
slip event in ~A.D. 1314, suggesting that the 2010 earthquake was not
an isolated event (Philibosian et al., 2012).
2.6. Sumatra subduction zone, northern segment
The great Mw 9.2 Sumatra earthquake of 26 December 2004 propagated ~1300 km along strike and produced a devastating tsunami. Singh
et al. (2008) imaged large, active thrust faults at the front of the accretionary wedge with seismic reection data. They also note the presence of aftershocks with thrust mechanisms consistent with the imaged faults.
These observations suggest that the 2004 earthquake may have propagated updip to the tip of the accretionary prism. This is compatible with geodetic data (Rhie et al., 2007), and supported by backprojections of
radiated energy, which indicate signicant energy release near the trench
for hundreds of kilometers along strike (Ishii et al., 2005). However, the
resolution of the slip distribution based on seismic energy release and
geodetic deformation is poor, and several alternative slip models t the
data reasonably well with primarily deeper slip patches (Ammon et al.,
2005; Chlieh et al., 2007; Pietrzak et al., 2007; Shearer and Brgmann,
2010). This is a common problem for offshore earthquakes far from
land. If signicant slip near the trench did occur, it could have produced
large uplift of the seaoor and contributed to the height of the tsunami.
Nearly a century prior to the 2004 Sumatran earthquake, a ~ M 7.8
earthquake occurred near the southern extent of the 2004 event in
1907 (Kanamori et al., 2010). This earthquake produced an extensive
tsunami that affected nearly 950 km of the Sumatran coast. Detailed investigations of historical seismograms has led to the conclusion that this
earthquake probably initiated on the subduction interface, and propagated up-dip into the shallow sediments, causing the large tsunami
(Kanamori et al., 2010).

rupture within the low-rigidity materials of an accretionary prism


(Newman and Okal, 1998; Polet and Kanamori, 2000). Thus, we infer
that the Java portion of the Sunda subduction zone, like Sumatra to
the north, may be capable of large ruptures in the frontal part of the
wedge.
2.8. Japan Trench
The 2011 Mw 9.0 Tohoku-Oki earthquake ruptured the plate interface between the Pacic and Okhotsk plates, east of the island of Honshu. Tremendous slip vertically displaced the sea bottom by up to
50 m, as measured by differential bathymetry, GPS measurements, underwater acoustic sounding, and sea-bottom pressure records of the
tsunami waves, demonstrating that rupture reached close to the Japan
trench (Avouac, 2011; Fujiwara et al., 2011; Sato et al., 2011). Indeed,
seismic reection proles pre- and post-earthquake directly image
slip at the trench axis during the earthquake (Kodaira et al., 2012),
and demonstrate that the dcollement underlying this prism slipped
in this event, with large coseismic slip reaching the toe of the wedge.
Should we have known prior to the Tohoku-Oki earthquake that this
subduction zone was capable of trench ruptures? Indeed, we should
have. The Mw 8.5 1896 Sanriku earthquake ruptured a nearby portion
of the trench and generated a tsunami with a maximum run-up of
25 m, despite only weak shaking being felt onshore (Tanioka and
Satake, 1996). Tsunami modeling suggests that slip was concentrated
near the trench (Tanioka and Satake, 1996). Repeated greater-thanexpected tsunamis with weak shaking support the hypothesis of multiple toe ruptures.
2.9. Kuril Trench
In addition to the two earthquakes in the southern part of the JapanKuril trench, two tsunami earthquakes occurred further north: a Ms 7.2
event in 1963, and a Ms 7.0 event in 1975 (Fukao, 1979; Pelayo and
Wiens, 1992). The Oct. 20 1963 earthquake occurred as an aftershock
of a larger, more down-dip earthquake on Oct. 13, and has been
interpreted as having a shallow origin (~ 5 km depth, Fukao, 1979;
~ 9 km depth, Pelayo and Wiens, 1992). The June 10, 1979 tsunami
earthquake occurred about 300 km southwest of the 1963 earthquake;
body wave inversion provides a best-t source depth of 5 km, which is
consistent with the location of the epicenter relative to the trench
(Pelayo and Wiens, 1992). Both of these events exhibited anomalously
slow rupture speeds, suggesting that they occurred within the weak
sediments of the accretionary prism (Pelayo and Wiens, 1992).
In 2006, a Mw 8.3 event ruptured just northeast of the 1963 earthquake region (Ammon et al., 2008; Lay et al., 2009). Although the tsunami was relatively limited in this event (mostly b1.2 m), this is in part
because of the lack of local tide gauge recordings. A nite source
model for the rupture suggests that at shallow depths (b25 km), a
250 km long segment of the subduction zone slipped an average of
4.3-6.5 m (Lay et al., 2009). This calculation constrains slip at the trench
to zero. It seems likely that this rupture extended to the trench.
2.10. Japan, Nankai Trough

2.7. Java subduction zone


The 17 July 2006 Mw 7.8 Java earthquake produced a tsunami with
local runup heights over 20 m and average tsunami heights of 5 to
7 m along 200 m of coastline (Ammon et al., 2006; Fritz et al., 2007).
Like the 2010 Mentawai earthquake, this qualies as a tsunami earthquake, due to the size of the tsunami relative to the earthquake magnitude (Kanamori, 1972). Also like the Mentawai earthquake, a rupture
towards the toe of the accretionary wedge could produce signicant uplift, and generate a tsunami of larger size than would be predicted by a
deeper rupture. This is consistent with the observation of low energy release and slow rupture, features that are inferred to be associated with

High-resolution seismic reection proles across the Nankai Trough


image both a series of shallowly dipping thrust faults in the frontal part
of the wedge and a large thrust to the hinterland termed a megasplay
that merges with the subduction interface at ~89 km depth (Park et al.,
2002; Moore et al., 2009). Cores were drilled through both the frontal
thrust at the toe of the accretionary prism and across the megasplay,
and analyzed for vitrinite reectance geothermometry (Sakaguchi
et al., 2011). It was determined that both fault zones underwent localized temperatures of more than 380 C, implying that frictional heating
occurred, likely due to coseismic slip. This evidence strongly suggests
that the frontal dcollement slipped coseismically at least once.

J. Hubbard et al. / Earth-Science Reviews 141 (2015) 4555

Although it is not possible to link this recorded heating event to a particular earthquake, both the Mw 8.1 1944 Tonankai and the Mw 8.1 1946
Nankaido earthquakes are potential candidates, as they produced
ground shaking and damaging tsunamis in this region (Kato and
Ando, 1997).

2.11. Solomon Islands


The 2007 Mw 8.1 Solomon Islands earthquake produced a large tsunami, with up to 12 m of runup (Chen et al., 2009; Furlong et al., 2009).
Inversion of geodetic data and surveys of uplifted coral reefs and other
coastal features suggests that there was up to 30 m of slip at the trench
(Chen et al., 2009). A few years later, the 2010 Mw 7.1 Solomon Islands
tsunami earthquake produced widespread coseismic subsidence within
20 km of the San Cristobal trench (Newman et al., 2011). Analysis of the
near-trench deformation, tsunami run-up, open-ocean wave height
data, and seismic records indicates that the earthquake occurred on
the shallow part of the low-angle dcollement in this region. The bestt slip model suggests that the maximum slip was close to the trench
and reached over 7 m (Newman et al., 2011).

2.12. Middle America megathrust


The September 2 1992 M 7.6 Nicaragua earthquake generated up to
10 m of tsunami runup, but only exhibited weak ground shaking. Extensive studies of this earthquake and its associated tsunami demonstrate
that it ruptured a 40-50 km wide fault plane extending from the trench
to depths of ~10 km (Satake, 1994; Lay and Bilek, 2007). Finite-fault inversion based on teleseismic P wave inversion shows slip down to
~ 20 km depth, but with large amounts within the upper 10 km (Ye
et al., 2013). In 2012, another earthquake ruptured the Middle
America megathrust, this time a Mw 7.3 in El Salvador. Both this event
and the 1992 earthquake exhibit low radiated energy for their seismic
moment and slow rupture velocities, characteristics of tsunami earthquakes (Kanamori and Kikuchi, 1993; Ye et al., 2013). Although the
2012 earthquake did not produce a tsunami due to its relatively low
magnitude, nite-fault inversion based on teleseismic P-wave inversion
indicates that most of the slip was within the upper 15 km. The inversion model of Ye et al. (2013) imposes zero slip at the trench, but
exhibits a steep slip gradient in the upper 5 km.
Kanamori and Kikuchi (1993) suggest that the shallow rupture in
1992 may have been made possible by the lack of a sedimentary
wedge offshore, resulting in different frictional properties than in
other wedges. Heesemann et al. (2009) propose that seamounts on
the subducting Cocos plate might alter the stress state and promote dynamic slip on a normally aseismic dcollement. However, our discussion
here demonstrates that such ruptures have occurred in many other subduction zones without these characteristics.

2.13. Alaskan Trench


The 1964 Mw 9.2 Great Alaskan earthquake ruptured the eastern segment of the Alaska-Aleutian Trench. Joint inversion of tsunami and geodetic data indicates that a large area of high slip concentrated over Prince
William Sound extended to the trench, with an average of ~11 m of slip
in the region above 15 km depth (Johnson et al., 1996). However,
the resolution of this slip model is limited. Slip certainly extended
as far out as Middleton Island, which lies about 8 km above the
megathrust: uplift in the earthquake produced the most recent of
six beach terraces visible on the island (Savage et al., 2014). Whether the slip continued to the trench or instead was diverted onto the
splay fault that is presumably responsible for the growth of this island (or both) is not known.

51

2.14. Peru
The Mw 7.6 Peru tsunami earthquake of Nov. 20, 1960, had a much
longer rupture duration than would have been otherwise expected for
an earthquake of this magnitude (Pelayo and Wiens, 1990), suggesting
that it occurred within the low-rigidity accretionary prism. Body waveform inversion indicates that it ruptured a fault plane dipping ~6 close
to the trench axis (Pelayo and Wiens, 1992). This earthquake occurred
in a portion of the Peru trench that previously had been mostly aseismic,
leading to the suggestion that it might not be capable of generating large
earthquakes (Nishenko, 1991). However, the 1960 event demonstrated
that shallow, tsunamigenic earthquakes can occur in this region, and
may occur in other, apparently aseismic regions as well.
The 1960 earthquake was followed by two other events with evidence of shallow slip on the Peruvian subduction zone: a Mw 7.4 tsunami earthquake in February 1996, which generated a large tsunami than
expected from its surface magnitude and exhibited a long rupture duration (Heinrich et al., 1998), and a Mw 8.5 earthquake in June 2001, for
which joint inversion of seismic and geodetic data suggest signicant
slip above 15 km depth, reaching close to the trench (Pritchard et al.,
2007).
2.15. Chile
The great 1960 Mw 9.5 Chile earthquake was the largest earthquake
ever recorded. Inversions for the slip distribution from ground deformation using a 3D fault model suggest that there was at least 1 m of slip at
the trench for a distance of 200 km, with 50 km of that distance
experiencing over 5 m of slip (Barrientos and Ward, 1990). In addition,
they note that substantial undetected offshore slip is likely, given the
large tsunami generated by the earthquake. Moreno et al. (2009) used
the same geodetic dataset combined with a 3D fault model to estimate
that over 200 km of the trench along strike slipped N 10 m, with over
50 km of that distance experiencing N 20 m of slip. A joint inversion of
geodetic and tsunami data for the 1960 earthquake estimates 13-21 m
of slip near the trench in the southern part of the source (Fujii and
Satake, 2013). These studies demonstrate the importance of considering
tsunami data in slip inversions, since geodetic data onshore cannot effectively constrain slip occurring at the trench.
2.16. Summary
The fteen regional examples that we describe include continentcontinent, continent-ocean, and ocean-ocean convergent settings, and
even oblique portions of transform margins; subduction zones with
old and young downgoing plates, with and without sediment deposition at the trench. They include fold-and-thrust belts that extend for
1000s of km and b100 km along strike, and systems that deform primarily through forethrusts, backthrusts, and combinations of the two.
They include very large and moderate-sized earthquake ruptures, and
ruptures that initiate at deep and shallow levels. Thus, it is not possible
to point to one common feature that makes rupture possible on the
shallow dcollement, implying that these faults may always have the
potential to be seismic and/or seismogenic.
3. How can weak decollements behave seismically?
Given that there is evidence for large, episodic slip at the tips of
dcollements, we must reevaluate the assumption that dcollements
are too weak to support stresses and store large strain energy. However, critical taper wedge mechanics implies that these are weak features, signicantly weaker than the surrounding rocks (Suppe,
2007). Reconciling these apparently conicting views requires us
to examine our assumptions about fault strength, earthquake dynamics and rock properties.

52

J. Hubbard et al. / Earth-Science Reviews 141 (2015) 4555

3.1. The state of stress within the wedge, and preferred slip planes at
geological time scales
We know that dcollements are weak because they form and slip at
unfavorable angles. However, we could turn this around and say instead
that these faults are oriented at unfavorable angles, and therefore the differential stress in the ambient rocks must be high in order for these faults
to slip. In fact, because we see thrust faults rising from dcollements, we
know that sometimes it is easier to break a more steeply dipping thrust
fault and slip on it than to slip on the dcollement surface. At other
times, however, it must be easier to slip on the dcollement; we know
this because we see slip at the toes.
How is it possible that both of these options are favorable at different
times? We can use a Mohr diagram, which describes the normal and
shear stresses acting on a plane of any orientation, and compares
them to failure criteria, to try to understand this problem (Fig. 4). We
propose a setting where minor changes in stress or rock properties
can cause slip on either the dcollement or on a thrust fault to be preferable at a given time.

However, we do not see random variations in slip on the dcollement


vs. thrusting. Instead, we typically observe in-sequence break-forward
propagation: the thrust faults towards the toe are more active, and
those in the hinterland have been abandoned (Morley, 1988). Thus, ruptures today are typically choosing to continue along the dcollement all
the way to the toe rather than to break upward along preexisting thrust
faults, even though they previously chose those branches. This pattern of
break-forward propagation can be understood by the fact that the principal stresses are rotated within the wedge due to the topographic gradient (Dahlen et al., 1984; Savage et al., 1985; Cubas et al., 2008; Mary et al.,
2013). The maximum stress direction dips gently in the direction of the
wedge toe (see 1 orientation on the right side of Fig. 4). This increases
the angle between the maximum stress orientation, 1, and the
dcollement. There is a lateral transition in the amount of stress rotation,
with 1 horizontal just beyond the toe of the wedge to tilted in the interior of the wedge. A larger angle between the 1 direction and the
dcollement dip is equivalent to a smaller in the Mohr diagram
(Fig. 4a). This makes slip on the dcollement more favorable. Thus, we
expect ruptures to follow the dcollement until they reach a point

Fig. 4. Mohr circle analysis of failure in a critical taper wedge. (A) Basic concept of a Mohr circle. Every point on the circle represents the normal () and shear () stresses acting on a plane
at an orientation that forms an angle with respect to the primary stress orientation 1. The size and location of the circle is dened by the largest (1) and smallest (3) principal stresses.
The black line represents the failure criterion of the material, with slope = the angle of internal friction, and the y-axis intercept for the failure criterion (here zero) represents the cohesion (here considered cohesionless). If the circle intersects the failure criterion, the material will fail on the plane represented by the point of intersection. (B) Mohr circle analysis of the
Alaskan wedge at the toe of the wedge and within the interior. Here we consider the potential for failure of the dcollement and a ramp rising from the dcollement. r and d represent the
angles of internal friction for the ramp and the detachment, respectively. A value of r = 0.64 was chosen to match the observed dips of the fault ramps (assumed to be the ideal orientation
for rupture). d can have a value as low as 0.035 without triggering slip on the dcollement at the toe, and must have a value below 0.187 to allow slip in the interior of the wedge. The
shaded blue area on the Mohr circles represents this range of possible d values. Note that these values would change in the presence of pore pressures or cohesion, but the principle would
remain the same. At the toe of the wedge, 1 is horizontal, and the ramp reaches failure, even though the ramp material has a higher coefcient of friction. In the interior, 1 dips slightly
towards the toe, rotating the planes represented on the Mohr circle clockwise. This causes the dcollement, rather than the ramp, to reach failure. The Mohr circle on the right shows lines
for both the unrotated and rotated 1 orientation, to better illustrate the effect of the rotation.

J. Hubbard et al. / Earth-Science Reviews 141 (2015) 4555

53

(usually near the toe) where the stress eld is less tilted, and at that point
to break upward along a thrust fault. In other words, the slip on the
ramps builds a topographic gradient that causes those ramps to shut
off, and the dcollement to propagate forward.

surface (Fig. 3). If slip propagates to the tip of the dcollement, we can
expect that it will then extend upward onto such a thrust fault and produce a band of high uplift at the toe of the wedge, resulting in larger
tsunamis.

3.2. Variations in strength through the earthquake cycle

5. Discussion and conclusions

There are additional mechanisms that could allow stress to accumulate on dcollements. In particular, the interseismic and coseismic
strengths of the dcollement may differ. In this case, the long-term record visible as the total deformation of the system may reect the weaker of the two. Low-temperature mechanical processes of weakening and
healing (abrasion, pulverization, and fusion) are not likely to be able to
dramatically affect effective friction (Biegel et al., 1992; Wang and
Scholz, 1994). However, at high slip velocities, frictional heating may
allow strong-weakening phenomena, such as melt-welts (Brown and
Fialko, 2012) and other ash-weakening effects (Rice, 2006) and pore
pressurization through thermal expansion (Ghabezloo and Sulem,
2009; Ferri et al., 2010) or mineral decomposition (Han et al., 2007).
Strong-weakening mechanisms are activated under specic conditions,
so we can expect different frictional resistance at subseismic and seismic slip speeds. For example, reaching sufciently high temperature
for melting or for substantial expansion of pore uids requires large
slip (Rice, 2006), so these strong-weakening effects are expected to
occur mostly during large earthquakes. But the complexity of fault rheology may also affect the interseismic period. For example, slow slip
events occurred near the hypocenter of the 2011 Mw 9.0 Tohoku-Oki
earthquake and the 2014 Mw 8.1 Iquique, Chile earthquake with significant overlap with the seismic rupture (Ito et al., 2013; Kato and
Nakagawa, 2014; Ruiz et al., 2014), and models of fault slip evolution
over many earthquake cycles suggest the possibility that the same segment may experience creep, slow slip, and seismic ruptures during different periods (Noda and Lapusta, 2013; Noda and Hori, 2014). This
indicates that the seismogenic potential of faults may not be fully
assessed during a short period of observation.

For many years, studies of subduction zones have assumed that the
frontal portions of accretionary prisms were too weak to rupture
coseismically (Byrne et al., 1988; Hyndman et al., 1997). Hazard studies
of fold-and-thrust belts on land have often suffered from the same assumption, based on the assumption that the shallow dcollement is
too weak to rupture in large earthquakes. We present a series of examples, both on-land and offshore, demonstrating that in many cases, shallow dcollements are capable of producing large, coseismic slip events
that rupture to the toes of the systems. Some of these events are associated with ruptures that initiate down-dip of the seismic front, while
others are limited to the frontal, shallow portion of the wedge
We suggest that this behavior is not limited to the examples described here, but rather is common to many (perhaps most) accretionary wedges and fold-and-thrust belts. Although many earthquakes in
subduction zones have been interpreted to have no slip at the tip of
the accretionary prism, this interpretation is typically driven by model
assumptions, rather than the data. In addition to the examples provided
here, there may be many other examples of similar earthquakes, where
known slip downdip obscures the need for slip at the toe (hidden tsunami earthquakes).
We do not characterize the regions and events described here as unusual, as they encompass a wide range of settings. However, we recognize that as far as reliably documented events go, these events are still
fairly limited. Whether this fact is due to a lack of data or a lack of events
is not clear, and accurately assessing this distinction may require the acquisition of new forms of data, including pre- and post- bathymetry surveys and seaoor geodesy (as has been used to observe deformation in
the 2011 Tohoku-Oki earthquake). The fraction of slip at the trench that
occurs in large seismogenic events is unknown, and likely varies between dcollements, and thus assessing the recurrence rates of such
earthquakes will in many cases be difcult, especially for subduction
zones, where paleoseismic evidence is lacking. As a result, it would be
safest to assume that trench-rupturing events can occur until we have
enough evidence to show that a particular fault system is behaving differently. Many attempts to assess seismic hazard for particular regions
have relied on extrapolating the likelihood of large earthquakes from
the occurrence of small ones, resulting in unexpected large earthquakes with striking underestimations (by two to three orders of magnitude) of expected fatalities (Kossobokov and Nekrasova, 2012; Wyss
et al., 2012). This study indicates that there is an urgent need for studies
that evaluate seismic and tsunami hazard in fold-and-thrust belts and
subduction zones around the world and allow for the possibility of shallow dcollement rupture.

4. Consequences of seismogenic decollements


Dcollements are some of the largest faults on Earth. However,
existing scaling relationships between fault area and earthquake magnitude (Wells and Coppersmith, 1994; Hanks and Bakun, 2008) may be
inappropriate for dcollements, as they are structurally different from
typical dip-slip or strike-slip faults. In addition, weak dcollements
and accretionary wedges exhibit much lower rigidity than more consolidated rock (Kanamori, 1972). Because earthquake moment is dened
as the product of the rigidity, slip, and area of a fault, lower rigidity
will result in lower earthquake magnitudes for a given amount of slip
and area (Fukao, 1979).
Such an alteration to the typical slipmagnitude relationship would
reduce expected seismic moments in regions where hazard is estimated
from slip, uplift amounts, or GPS shortening. Initially, one would expect
this alteration to reduce seismic hazard. However, because the slip
would be expected to occur directly beneath large regions at shallow
depths, the resulting ground shaking levels might be as or more hazardous than that predicted from higher seismic moments on more distant
faults.
In contrast, if seismic hazard is estimated from earthquake moment,
we may see earthquakes with higher slip than expected based on seismic records. This would help explain tsunami earthquakes, as even
allowing slip to propagate to the tips of accretionary prisms is not always sufcient to explain resulting tsunami heights (e.g., Hill et al.,
2012).
This discrepancy between earthquake magnitude and slip can also
be partially solved by structural analysis. Seismic reection imaging of
accretionary prisms demonstrates that dcollement tips frequently terminate into more steeply dipping thrust faults that break towards the

Acknowledgements
We thank our two reviewers, Jeff Freymueller and Thorne Lay, for
their constructive comments. This research was supported by the National Research Foundation of Singapore under the NRF Fellowship
scheme (National Research Fellow Award No. NRF-NRFF2013-06) and
by the EOS, the National Research Foundation of Singapore and the
Singapore Ministry of Education under the Research Centres of Excellence initiative. This is EOS paper number 76.
References
Adam, J., Klaeschen, D., Kukowski, N., Flueh, E., 2004. Upward delamination of Cascadia
Basin sediment inll with landward frontal accretion thrusting caused by rapid glacial
age material ux. Tectonics 23 (TC3009). http://dx.doi.org/10.1029/2002TC001475.

54

J. Hubbard et al. / Earth-Science Reviews 141 (2015) 4555

Ader, T., Avouac, J.-P., Liu-Zeng, J., Lyon-Caen, H., Bollinger, L., Galetzka, J., Genrich, J.,
Thomas, M., Chanard, K., Sapkota, S.N., Rajaure, S., Shrestha, P., Ding, L., Flouzat, M.,
2012. Convergence rate across the Nepal Himalaya and interseismic coupling on
the Main Himalayan Thrust: Implications for seismic hazard. J. Geophys. Res. 117
(B04403). http://dx.doi.org/10.1029/2011JB009071.
Ammon, C.J., Ji, C., Thio, H.-K., Robinson, D., Ni, S., Hjorleifsdottir, V., Kanamori, H., Lay, T.,
Das, S., Helmberger, D., Ichinose, G., Polet, J., Wald, D., 2005. Rupture process of the
2004 Sumatra-Andaman earthquake. Science 308, 11331139.
Ammon, C.J., Kanamori, H., Lay, T., Velasco, A.A., 2006. The 17 July 2006 Java tsunami earthquake. Geophys. Res. Lett. 33 (L24308). http://dx.doi.org/10.1029/
2006GL028005.
Ammon, C.J., Kanamori, H., Lay, T., 2008. A great earthquake doublet and seismic stress
transfer cycle in the central Kuril islands. Nature 451. http://dx.doi.org/10.1038/
nature06521.
Avouac, J.-P., 2011. Earthquakes: The lessons of Tohoku-Oki. Nature 475. http://dx.doi.
org/10.1038/nature10265.
Barbot, S., Agram, P., De Michele, M., 2013. Change of apparent segmentation of the San
Andreas fault around Parkeld from space geodetic observations across multiple periods. J. Geophys. Res. 118, 63116327.
Barrientos, S.E., Ward, S.N., 1990. The 1960 Chile earthquake inversion for slip distribution from surface deformation. Geophys. J. Int. 103 (3), 589598.
Behrmann, J.H., Brown, K., Moore, J.C., Mascle, A., Taylor, E., Alvarez, F., Andreiff, P., Barnes,
R., Beck, C., 1988. Evolution of structures and fabrics in the Barbados accretionary
prism. Insights from Leg 110 of the Ocean Drilling Program. J. Struct. Geol. 10 (6),
577591.
Betinelli, P., Avouac, J.-P., Flouzat, M., Jouanne, F., Bollinger, L., Willis, P., Chitrakar, G.,
2006. Plate motion of India and interseismic strain in the Nepal Himalaya from GPS
and DORIS measurements. J. Geod. 80, 567589.
Biegel, R.L., Wang, W., Scholz, C.H., Boitnott, G.N., 1992. Micromechanics of rock friction. 1.
Effects of surface roughness on initial friction and slip hardening in Westerly granite.
J. Geophys. Res. 97 (B6), 89518964.
Bilham, R., Larson, K., Freymueller, J., 1997. GPS measurements of present-day convergence across the Nepal Himalaya. Nature 386, 6164.
Bilotti, F., Shaw, J.H., 2005. Deep-water Niger Delta fold and thrust belt modeled as a
critical-taper wedge: The inuence of elevated basal uid pressure on structural
styles. AAPG Bull. 89 (11), 14751491.
Bollinger, L., Sapkota, S.N., Tapponnier, P., Klinger, Y., Rizza, M., Van der Woerd, J., Tiwari,
D.R., Pandey, R., Bitri, A., Bes de Berc, S., 2014. J. Geophys. Res. http://dx.doi.org/10.
1002/2014JB010970.
Brooks, B.A., Arrowsmith, R., Echalar, A., Ericksen, T.L., Weiss, J.R., Ahlgren, K., Bevis, M.G.,
Whipple, K.X., 2011a. Great earthquakes and orogenic wedge front processes in the
Bolivian backarc. Am. Geophys. U. Fall Meeting 2011 (abstract #T14C-03).
Brooks, B.A., Bevis, M., Whipple, K., Arrowsmith, J.R., Foster, J., Zapata, T., Kendrick, E.,
Minaya, E., Echalar, A., Blanco, M., Euillades, P., Sandoval, M., Smalley Jr., R.J., 2011b.
Orogenic-wedge deformation and potential for great earthquakes in the central Andean backarc. Nat. Geosci. 4, 380383. http://dx.doi.org/10.1038/ngeo1143.
Brown, K.M., Fialko, Y., 2012. Melt welt mechanism of extreme weakening of gabbro at seismic slip rates. Nature 488, 638642. http://dx.doi.org/10.1038/
nature11370.
Burbidge, D.R., Braun, J., 2002. Numerical models of the evolution of accretionary wedges
and fold-and-thrust belts using the distinct-element method. Geophys. J. Int. 148 (3),
542561.
Byrne, D.E., Davis, D.M., Sykes, L.R., 1988. Loci and maximum size of thrust earthquakes
and the mechanics of the shallow region of subduction zones. Tectonics 7 (4),
833857.
Chen, T., Newman, A.V., Feng, L., Fritz, H.M., 2009. Slip distribution from the 1 April 2007
Solomon Islands earthquake: A unique image of near-trench rupture. Geophys. Res.
Lett. 36 (L16307). http://dx.doi.org/10.1029/2009GL039496.
Chlieh, M., Avouac, J.-P., Hjorleifsdottir, V., Song, A.T.-R., Ji, C., Sieh, K., Sladen, A.,
Hebert, H., Prawirodirdjo, L., Bock, Y., Galetzka, J., 2007. Coseismic slip and
afterslip of the Great Mw 9.15 Sumatra-Andaman earthquake of 2004. Bull.
Seismol. Soc. Am. 97 (1A).
Chlieh, M., Avouac, J.-P., Sieh, K., Natawidjaja, D.H., Galetzka, J., 2008. Heterogeneous coupling of the Sumatran megathrust constrained by geodetic and paleogeodetic measurements. J. Geophys. Res. 113, B05305.
Cubas, N., Leroy, Y.M., Maillot, B., 2008. Prediction of thrusting sequences in accretionary
wedges. J. Geophys. Res. 113, B12412.
Cubas, N., Avouac, J.-P., Souloumiac, P., Leroy, Y., 2013. Megathrust friction determined
from mechanical analysis of the forearc in the Maule earthquake area. Earth Planet.
Sci. Lett. 381, 92103.
Dahlen, F.A., 1990. Critical taper model of fold-and-thrust belts and accretionary wedges.
Annu. Rev. Earth Planet. Sci. 18, 5599.
Dahlen, F.A., Suppe, J., Davis, D., 1984. Mechanics of fold-and-thrust belts and accretionary
wedges: Cohesive Coulomb theory. J. Geophys. Res. 88 (B2), 11531172.
Davis, D., Suppe, J., Dahlen, F.A., 1983. Mechanics of fold-and-thrust belts and accretionary
wedges. J. Geophys. Res. 88 (B2), 11531172.
Faulkner, D.R., Mitchell, T.M., Behnsen, J., Hirose, T., Shimamoto, T., 2011. Geophys. Res.
Lett. 38 (L18303). http://dx.doi.org/10.1029/2011GL048552.
Ferri, F., Di Toro, G., Hirose, T., Shimamoto, T., 2010. Evidence of thermal pressurization in high-velocity friction experiments on smectite-rich gouges. Terra
Nova 22 (5), 347353. http://dx.doi.org/10.1111/j.1365-3121.2010.00955.x.
Fritz, H.M., et al., 2007. Extreme runup from the 17 July 2006 Java tsunami. Geophys. Res.
Lett. 34, L12602. http://dx.doi.org/10.1029/2007GL029404.
Fujii, Y., Satake, K., 2013. Slip distribution and seismic moment of the 2010 and 1960
Chilean earthquakes inferred from tsunami waveforms and coastal geodetic data.
Pure Appl. Geophys. 170, 14931509.

Fujiwara, T., Kodaira, S., No, T., Yuka, Kaiho, Takaashi, N., Kaneda, Y., 2011. The 2011
Tohoku-Oki earthquake: Displacement reaching the trench axis. Science 334, 12401240.
Fukao, Y., 1979. Tsunami earthquakes and subduction processes near deep-sea trenches.
J. Geophys. Res. 84 (B5), 23032314.
Furlong, K.P., Lay, T., Ammon, C.J., 2009. A great earthquake rupture across a rapidly
evolving three-plate boundary. Science 324, 226229.
Ghabezloo, S., Sulem, J., 2009. Stress dependent thermal pressurization of a uidsaturated rock. Rock Mech. Rock. Eng. 42 (1), 124. http://dx.doi.org/10.1007/
s00603-008-0165-z.
Graveleau, F., Malavieille, J., Dominguez, S., 2012. Experimental modeling of orogenic
wedge: A review. Tectonophysics 538. http://dx.doi.org/10.1015/j.tecto.2012.01.
027.
Han, R.H., Shimamoto, T., Ando, J.I., Ree, J.H., 2007. Seismic slip record in carbonatebearing fault zones: An insight from high-velocity friction experiments on siderite
gouge. Geology 35 (12), 11311134.
Hanks, T.C., Bakun, W.H., 2008. M-logA observations for recent large earthquakes. Bull.
Seismol. Soc. Am. 98 (1), 490494.
Heesemann, M., Grevemeyer, I., Villinger, H., 2009. Thermal constraints on the frictional
conditions of the nucleation and rupture area of the 1992 Nicaragua tsunami earthquake. Geophys. J. Int. 179, 12651278.
Heinrich, P., Schindele, F., Guibourg, S., Ihml, P.F., 1998. Modeling of the February 1996
Peruvian tsunami. Geophys. Res. Lett. 25 (14), 26872690.
Hill, E.M., Borrero, J.C., Huang, Z., Qiu, Q., Banerjee, P., Natawidjaja, D.H., Eolsegui, P.,
Fritz, H.M., Suwargadi, B.W., Pranantyo, I.R., Li, L., Macpherson, K.A., Skanavis, V.,
Synolakis, C.E., Sieh, K., 2012. The 2010 Mw 7.8 Mentawai earthquake: Very shallow source of a rare tsunami earthquake determined from tsunami eld survey
and near-eld GPS data. J. Geophys. Res. 117 (B06402). http://dx.doi.org/10.
1029/2012JB009159.
Hubbard, J., Shaw, J.H., Klinger, Y., 2010. Structural setting of the 2008 Mw7.9 Wenchuan,
China, earthquake. Bull. Seismol. Soc. Am. 100, 27132735.
Hubbard, J., Shaw, J.H., Dolan, J., Pratt, T.L., McAuliffe, L., Rockwell, T.K., 2014. Structure and
seismic hazard of the Ventura Avenue anticline and Ventura fault, California: Prospect
for large, multi-segment ruptures in the Western Transverse Ranges. Bull. Seismol.
Soc. Am. 104 (3), 10701087.
Hyndman, R.D., Yamano, M., Oleskevich, D.A., 1997. The seismogenic zone of subduction
thrust faults. Island Arc 6, 244260.
Ishii, M., Shearer, P.M., Houston, H., Vidale, J.E., 2005. Extent, duration and speed of the
2004 Sumatra-Andaman earthquake imaged by the Hi-Net array. Nature 435
(7044), 933936. http://dx.doi.org/10.1038/nature03675.
Ito, Y., Hino, R., Kido, M., Fujimoto, H., Osada, Y., Inazu, D., Ohta, Y., Iinuma, T., Ohzono, M.,
Miuraand, S., Mishina, M., Suzuki, K., Tsuji, T., Ashi, J., 2013. Episodic slow slip events
in the Japan subduction zone before the 2011 Tohoku-Oki earthquake. Tectonophysics
600, 1426. http://dx.doi.org/10.1016/j.tecto.2012.08.022.
Jackson, M., Bilham, R., 1994. Constraints on Himalayan deformation inferred from vertical velocity elds in Nepal and Tibet. J. Geophys. Res. 99 (B7), 13,89713,912.
Johnson, J.M., Satake, K., Holdahl, S.R., Sauber, J., 1996. The 1964 Prince William Sound
earthquake: Joint inversion of tsunami and geodetic data. J. Geophys. Res. 101,
523532.
Johnson, K.M., Hsu, Y.J., Segall, P., Yu, S.B., 2001. Fault geometry and slip distribution of the
1999 Chi-Chi, Taiwan earthquake imaged from inversion of GPS data. Geophys. Res.
Lett. 28 (11), 22852288.
Jouanne, F., Mugnier, J., Gamond, J., Le Fort, P., Pandey, M., Bollinger, L., Flouzat, M.,
Avouac, J.-P., 2004. Current shortening across the Himalayas of Nepal. Geophys. J.
Int. 157, 114.
Kanamori, H., 1972. Mechanism of tsunami earthquakes. Phys. Earth Planet. Inter. 6,
346359.
Kanamori, H., Kikuchi, M., 1993. The 1992 Nicaragua earthquake: A slow tsunami earthquake associated with subducted sediments. Nature 361, 714716.
Kanamori, H., Rivera, L., Lee, W.H.K., 2010. Historical seismograms for unraveling a mysterious earthquake: The 1907 Sumatra Earthquake. Geophys. J. Int. http://dx.doi.
org/10.1111/j.1365-246X.2010.04731.x.
Kato, T., Ando, M., 1997. Source mechanisms of the 1944 Tonankai and 1946 Nankaido
earthquakes: Spatial heterogeneity of rise times. Geophys. Res. Lett. 24 (16),
20552058.
Kato, A., Nakagawa, S., 2014. Multiple slow-slip events during a foreshock sequence of the
2014 Iquique, Chile M-w 8.1 earthquake. Geophys. Res. Lett. 41 (15), 54205427.
http://dx.doi.org/10.1002/2014GL061138.
Kodaira, S., No, T., Nakamura, Y., Fujiwara, T., Kaiho, Y., Miura, S., Takahashi, N., Kaneda, Y.,
Taira, A., 2012. Coseismic fault rupture at the trench axis during the 2011 Tohoku-Oki
earthquake. Nat. Geosci. 5. http://dx.doi.org/10.138/NGEO1547.
Kossobokov, V.G., Nekrasova, A.K., 2012. Global seismic hazard assessment program maps
are erroneous. Seism. Instrum. 48 (2), 162170.
Kozdon, J.E., Dunham, E.M., 2013. Rupture to the trench: Dynamic rupture simulations of the 11 March 2011 Tohoku earthquake. Bull. Seismol. Soc. Am. 103
(2B), 12751289.
Larson, K.M., Brgmann, R., Bilham, R., Freymueller, J.T., 1999. Kinematics of the IndiaEurasia collision zone from GPS measurements. J. Geophys. Res. 104, 10771093.
Lav, J., Avouac, J.P., 2000. Active folding of uvial terraces across the Siwaliks Hills,
Himalayas of central Nepal. J. Geophys. Res. 105 (B3), 57355770.
Lay, T., Bilek, S.L., 2007. Anomalous earthquake ruptures at shallow depths on subduction
zone megathrusts. In: Dixon, T., Moore, C. (Eds.), The Seismogenic Zone of Subduction
Thrust Faults. Columbia University Press, pp. 476511.
Lay, T., Kanamori, H., Ammon, C.J., Hutko, A.R., Furlong, K., Rivera, L., 2009. The 2006-2007
Kuril Islands great earthquake sequence. J. Geophys. Res. 114 (B11308). http://dx.doi.
org/10.1029/2008JB006280.

J. Hubbard et al. / Earth-Science Reviews 141 (2015) 4555


Lee, S.-J., Ma, K.-F., 1999. Rupture process of the 1999 Chi-Chi, Taiwan, earthquake
from the inversion of teleseismic data. Terr. Atmos. Ocean. Sci. 11 (3),
591608.
Loveless, J.P., Meade, B.J., 2009. Geodetic mapping of plate motions, slip rates, and
partitioning of deformation in Japan. J. Geophys. Res. 115 (B02410).
Loveless, J.P., Meade, B.J., 2011. Spatial correlation of interseismic coupling and coseismic
rupture extent of the 2011 Mw=9.0 Tohoku-oki earthquake. Geophys. Res. Lett. 38
(L17306). http://dx.doi.org/10.1029/2011GL048561.
Ma, K.F., Song, T.R.A., Lee, S.J., Wu, H.I., 2000. Spatial slip distribution of the September 20,
1999, Chi-Chi, Taiwan, earthquake (Mw 7.6): inverted from teleseismic data.
Geophys. Res. Lett. 27, 34173420.
Ma, K.F., Brodsky, E.E., Mori, J., Ji, C., Song, T.R.A., 2003. Evidence for fault lubrication during the 1999 Chi-Chi, Taiwan, earthquake (Mw7.6). Geophys. Res. Lett. 30 (5), 1244.
http://dx.doi.org/10.1029/2002GL015380.
Malaveille, J., 2010. Impact of erosion, sedimentation, and structural heritage on the structure and kinematics of orogenic wedges: Analog models and case studies. GSA Today
20 (1). http://dx.doi.org/10.1130/GSATG48A.1.
Marone, C., Scholz, C.H., 1988. The depth of seismic faulting and the upper transition from
stable to unstable slip regimes. Geophys. Res. Lett. 15 (6), 621624.
Mary, B.C.L., Maillot, B., Leroy, Y.M., 2013. Deterministic chaos in frictional wedges revealed by convergence analysis. Int. J. Numer. Anal. Methods Geomech. http://dx.
doi.org/10.1002/nag.2177.
McCaffrey, R., 2008. Global frequency of magnitude 9 earthquakes. Geology 36 (3),
263266. http://dx.doi.org/10.1130/G24402A.1.
Meade, B.J., Loveless, J.P., 2009. Predicting the geodetic signature of Mw 8 slow
slip events. Geophys. Res. Lett. 36 (L01306). http://dx.doi.org/10.190/
2008GL036364.
Moore, G.F., et al., 2009. Structural and seismic stratigraphic framework of the
NanTroSEIZE Stage 1 transect. NanTroSEIZE Stage 1: Investigations of Seismogenesis,
Nankai Trough, Japan, Proc. Integr. Ocean Drill. Prog http://dx.doi.org/10.2204/iodp.
proc.314315316.102.2009 (314/315/316).
Moreno, M.S., Bolte, J., Klotz, J., Melnick, D., 2009. Impact of megathrust geometry on inversion of coseismic slip from geodetic data: Application to the 1960 Chile earthquake. Geophys. Res. Lett. 36 (L16310).
Morley, C.K., 1988. Out-of-sequence thrusts. Tectonics 7 (3), 539561.
Morley, C.K., King, R., Hillis, R., Tingay, M., Backe, G., 2011. Deepwater fold and thrust belt
classication, tectonics, structure and hydrocarbon prospectivity: A review. Earth Sci.
Rev. 104, 4191.
Newman, A.V., Okal, E.A., 1998. Moderately slow character of the 17 July 1998 Sandaun
earthquake as studied by teleseismic energy estimates. Eos Trans. AGU 79 (Fall
Meet. Suppl.), F564.
Newman, A.V., Feng, L., Fritz, H.M., Lifton, Z.M., Kalligeris, N., Wei, Y., 2011. The energetic 2010 Mw 7.1 Solomon Islands tsunami earthquake. Geophys. J. Int. 186,
775781.
Nishenko, S.P., 1991. Circum-Pacic seismic potential, 1989-1999. Pure Appl. Geophys.
135 (2), 169259.
Noda, H., Hori, T., 2014. Under what circumstances does a seismogenic patch produce
aseismic transients in the later interseismic period? Geophys. Res. Lett. http://dx.
doi.org/10.1002/2014GL061676.
Noda, H., Lapusta, N., 2013. Stable creeping fault segments can become destructive
as a result of dynamic weakening. Nature 493. http://dx.doi.org/10.1038/
nature11703.
Oleskevich, D.A., Hyndman, R.D., Wang, K., 1999. The updip and downdip limits to
great subduction earthquakes: Thermal and structural models of Cascadia,
south Alaska, SW Japan, and Chile. J. Geophys. Res. 104, 14,96514,991.
Pacheco, J.F., Sykes, L.R., Scholtz, C.H., 1993. Nature of seismic coupling along simple plate
boundaries of the subduction type. J. Geophys. Res. 98, 14,13314,159.
Pandey, M.R., Tandukar, R.P., Avouac, J.P., Lav, J., Massot, J.P., 1995. Interseismic strain accumulation on the Himalayan Crustal Ramp (Nepal). Geophys. Res. Lett. 22 (7),
751754.
Park, J.-O., Tsuru, T., Kodaira, S., Cummins, P.R., Kaneda, Y., 2002. Splay fault branching
along the Nankai subduction zone. Science 297, 11571160.
Pelayo, A.M., Wiens, D.A., 1990. The November 20, 1960 Peru tsunami earthquake: source
mechanism of a slow event. Geophys. Res. Lett. 17 (6), 661664.
Pelayo, A.M., Wiens, D.A., 1992. Tsunami earthquakes: Slow thrust-faulting events in the
accretionary wedge. J. Geophys. Res. 97 (B11), 15,32115,337.
Philibosian, B., Sieh, K., Natawidjaja, S.H., Chiang, H.-W., Shen, C.-C., Suwargadi, B.W., Hill,
E.M., Edwards, R.L., 2012. An ancient shallow slip event on the Mentawai segment of
the Sunda megathrust, Sumatra. J. Geophys. Res. 117 (B05401). http://dx.doi.org/10.
1029/2011JB009075.
Pietrzak, J., Socquet, A., Ham, D., Simons, W., Vigny, C., Labeur, R.J., Schrama, E., Stelling, G.,
Vatvani, D., 2007. Dening the source region of the Indian Ocean Tsunami from GPS,
altimeters, tide gauges and tsunami models. Earth Planet. Sci. Lett. 261, 4964.
Polet, J., Kanamori, H., 2000. Shallow subduction zone earthquakes and their
tsunamigenic potential. Geophys. J. Int. 142, 684702.
Pritchard, M.E., Norabuena, E.O., Ji, C., Boroschek, R., Comte, D., Simons, M., Dixon, T.H.,
Rosen, P.A., 2007. Geodetic, teleseismic, and strong motion constraints on slip from
recent southern Peru subduction zone earthquakes. J. Geophys. Res. 112 (B03307).
http://dx.doi.org/10.1029/2006JB004294.
Ranero, C.R., von Huene, R., Flueh, E., Duarte, M., Baca, D., McIntosh, K., 2000. A
cross section of the convergent Pacic margin of Nicaragua. Tectonics 19 (2),
335357.
Rhie, J., Dreger, D., Brgmann, R., Romanowicz, B., 2007. Slip of the 2004 SumatraAndaman earthquake from joint inversion of long-period global seismic waveforms
and GPS static offsets. Bull. Seismol. Soc. Am. 97, S115S127.

55

Rice, J.R., 2006. Heating and weakening of faults during earthquake slip. J. Geophys. Res.
111 (B5). http://dx.doi.org/10.1029/2005JB004006.
Rockwell, T.K., 2011. Large co-seismic uplift of coastal terraces across the Ventura
Avenue anticline: Implications for the size of earthquakes and the potential for
tsunami generation, Plenary talk. SCEC Annual Meeting (12 Sept).
Rousset, B., Barbot, S., Avouac, J.-P., Hsu, Y.-J., 2012. Postseismic deformation following the
1999 Chi-Chi earthquake, Taiwan: Implication for the lower-crust rheology. J.
Geophys. Res. 117, B12405.
Rubin, A.M., Gillard, D., Got, J.-L., 1999. Streaks of microearthquakes along creeping faults.
Nature 400, 635641.
Ruiz, S., Metois, M., Fuenzalida, A., Ruiz, J., Leyton, F., Grandin, R., Vigny, C., Madariaga, R.,
Campos, J., 2014. Intense foreshocks and a slow slip event preceded the 2014 Iquique
Mw 8.1 earthquake. Science 345 (6201), 11651169. http://dx.doi.org/10.1126/
science.1256074.
Sakaguchi, A., Chester, F., Curewitz, D., Fabbri, O., Goldsby, D., Kimura, G., Li, C.-F., Masaki,
Y., Screaton, E.J., Tsutsumi, A., Ujiie, K., Yamaguchi, A., 2011. Seismic slip propagation
to the updip end of plate boundary subduction interface faults: Vitrinite reectance
geothermometry on Integrated Ocean Drilling Program NanTro SEIZE cores. Geology
39 (4), 395398.
Sapkota, S.N., Bollinger, L., Klinger, Y., Tapponnier, P., Gaudemer, Y., Tiwari, D., 2013. Primary surface ruptures of the great Himalayan earthquakes in 1934 and 1255. Nat.
Geosci. 6 (1), 7176.
Satake, K., 1994. Mechanics of the 1992 Nicaragua tsunami earthquake. Geophys. Res.
Lett. 21, 25192522.
Sato, M., Ishikawa, T., Ujihara, N., Yoshida, S., Fujita, M., Mochizuki, M., Asada, A., 2011.
Displacement above the hypocenter of the 2011 Tohoku-Oki earthquake. Science
http://dx.doi.org/10.1126/science.1207401.
Savage, W.Z., Swolfs, H.S., Powers, P.S., 1985. Gravitational stresses in long symmetric
ridges and valleys. Int. J. Rock Mech. Min. Sci. 22 (5), 291302.
Savage, J.C., Plafker, G., Svarc, J.L., Lisowski, M., 2014. Continuous uplift near the seaward
edge of the Prince William Sound megathrust: Middleton Island, Alaska.
Geophys. J. Int. 119, 60676079. http://dx.doi.org/10.1002/2014JB011127.
Schelling, D., Arita, K., 1991. Thrust tectonics, crustal shortening, and the structure of the
far-Eastern Nepal Himalaya. Tectonics 10 (5), 851862.
Shaw, J.H., Suppe, J., 1994. Active faulting and growth folding in the eastern Santa Barbara
Channel, California. Geol. Soc. Am. Bull. 106, 607626.
Shearer, P., Brgmann, R., 2010. Lessons learned from the 2004 Sumatra-Andaman
megathrust rupture. Annu. Rev. Earth Planet. Sci. 38, 103131.
Shulgin, A., Kopp, H., Mueller, C., Planert, L., Lueschen, E., Flueh, E.R., Djajadihardja,
Y., 2011. Structural architecture of oceanic plateau subduction offshore Eastern Java and the potential implications for geohazards. Geophys. J. Int. 184,
1228.
Singh, S.C., Carton, H., Tapponnier, P., Hananto, N.D., Chauhan, A.P.S., Hartoyo, D., Bayly,
M., Moeljopranoto, S., Bunting, T., Christie, P., Lubis, H., Martin, J., 2008. Seismic evidence for broken oceanic crust in the 2004 Sumatra earthquake epicentral region.
Nat. Geosci. 1, 777781.
Singh, S.C., Hananto, N., Mukti, M., Permana, H., Djajadihardja, Y., Harjono, H., 2011. Seismic images of the megathrust rupture during the 25th October 2010 Pagai earthquake, SW Sumatra: Frontal rupture and large tsunami. Geophys. Res. Lett. 38
(L16313). http://dx.doi.org/10.1029/2011GL048935.
Strayer, L.M., Hudleston, P.J., Lorig, L.J., 2001. A numerical model of deformation and
uid-ow in an evolving thrust wedge. Tectonophysics 335 (12), 121145.
Suppe, J., 2007. Absolute fault and crustal strength from wedge tapers. GSA Bull. 35 (12),
11271130.
Tanioka, Y., Satake, K., 1996. Tsunami generation by horizontal displacement of ocean
bottom. Geophys. Res. Lett. 23 (8), 861864.
Uba, C.E., Kley, J., Strecker, M.R., Schmitt, A.K., 2009. Unsteady evolution of the
Bolivian Subandean thrust belt: The role of enhanced erosion and clastic
wedge progradation. Earth Planet. Sci. Lett. 281, 134146.
Venkataraman, A., Kanamori, H., 2004. Observational constraints on the fracture energy of
subduction zone earthquakes. J. Geophys. Res. 109, B05302.
Wang, W., Scholz, C.H., 1994. Micromechanics of the velocity and normal stress dependence of rock friction. Pure Appl. Geophys. 143 (1/2/3).
Wells, D.L., Coppersmith, K.J., 1994. New empirical relationship among magnitude, rupture length, rupture width, and surface displacement. Bull. Seismol. Soc. Am. 84 (4),
9741002.
Wyss, M., Nekrasova, A., Kossobokov, V., 2012. Errors in expected human losses due to incorrect seismic hazard estimates. Nat. Hazards 62, 927935. http://dx.doi.org/10.
1007/s11069-012-0125-5.
Ye, S., Flueh, E.R., Klaeschen, D., von Huene, R., 1997. Crustal structure along the EDGE
transect beneath the Kodiak shelf off Alaska derived from OBH seismic refraction
data. Geophys. J. Int. 130, 283302.
Ye, L., Lay, T., Kanamori, H., 2013. Large earthquake rupture process variations on
the Middle America megathrust. Earth Planet. Sci. Lett. 381, 147155.
Yue, L.-F., Suppe, J., Hung, J.-H., 2005. Structural geology of a classic thrust belt earthquake: the 1999 Chi-Chi earthquake, Taiwan (Mw = 7.6). J. Struct. Geol. 27,
20582083.
Yue, H., Lay, T., Rivera, L., Bai, Y., Yamazaki, Y., Fai Cheung, K., Hill, E.M., Sieh, K.,
Kongko, W., Muhari, A., 2014. Rupture process of the 2010 Mw 7.8 Mentawai
tsunami earthquake from joint inversion of near-eld hr-GPS and teleseismic
body wave recordings constrained by tsunami observations. J. Geophys. Res.
119. http://dx.doi.org/10.1002/2014JB011082.

You might also like