Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

SRL Early Edition


E

Coseismic Slip in the 2016 M w 7.8 Ecuador


Earthquake Imaged from Sentinel-1A Radar
Interferometry
by Ping He, Eric A. Hetland, Qi Wang, Kaihua Ding, Yangmao Wen, and
Rong Zou
ABSTRACT
The M w 7.8 Ecuador earthquake on 16 April 2016 is the sixth 1906 (M w 8.8) Colombia–Ecuador earthquake, which is the
earthquake larger than M w 7 to rupture the subduction mega- largest event documented offshore Ecuador and to date is the
thrust between the Nazca and South American plates since the seventh largest earthquake worldwide. Following the 1906
M w 8.8 Colombia–Ecuador earthquake in 1906. We use In- earthquake, there have been several other significant mega-
terferometric Synthetic Aperture Radar (InSAR) images from thrust earthquakes offshore Ecuador, including the 1942
Sentinel-1A to determine coseismic surface displacements (M w 7.8), 1958 (M w 7.7), 1979 (M w 8.2), and 1998 (M w 7.1),
associated with this earthquake. The interferograms exhibit along with the recent earthquake in 2016 (M w 7.8; Fig. 1). All
a relatively simple pattern of deformation, with maximum dis- of these earthquakes in the last century ruptured the mega-
placements of 0.7 and 0.3 m on the descending and ascending thrust within a 500-km-long seismic belt between latitude
images, respectively. We invert the interferograms for both rup- 3° N and 2° S (e.g., Swenson and Beck, 1996; Chlieh et al.,
ture geometry and slip distribution in the earthquake. We find 2014). Although each of the more recent earthquakes was
that the data are best described by slip on a fault dipping 17° to possibly due to failure of isolated asperities, the 1906 earth-
the east and that these InSAR data cannot uniquely constrain quake most likely involved rupture of multiple asperities (e.g.,
the strike. The maximum inferred slip is just over 2.5 m at Swenson and Beck, 1996; Chlieh et al., 2014). In this article,
about 20 km depth, with the main slip in the depth range of we focus on the recent megathrust earthquake offshore of
about 10–25 km. The geodetic moment of our preferred slip Ecuador, with a goal of placing this earthquake in the context
model is 7:15 × 1020 N·m, equivalent to M w 7.87. Our results of the past earthquakes offshore Ecuador and the 1906
suggest that there is little, if any, partitioning of the oblique Colombia–Ecuador earthquake.
plate convergence. The 2016 Ecuador earthquake is coincident On 16 April 2016 at 23:58:37 UTC, an M w 7.8 thrust
with the location of the 1942 M w 7.8 earthquake, with both earthquake occurred offshore of Ecuador. A maximum modi-
earthquakes most likely rupturing an asperity that also failed in fied Mercalli intensity (MMI) of VIII was reported in the cities
the 1906 M w 8.8 earthquake. of Esmeraldas and Pedernales, ∼130 and 170 km, respectively,
from the capital Quito, Ecuador (U.S. Geological Survey
[USGS]; see Data and Resources), causing over 450 fatalities
Electronic Supplement: Recorded and synthetic interferograms, (emergency events database [EM-DAT]; see Data and Resour-
figures of trade-off between fault parameters, checkerboard ces). During the first two months, the mainshock was followed
test, and coseismic slip model. by a total of 138 aftershocks with M w ≥ 3 (the largest M w 6.7)
within a nearly north–south-trending zone of 150 km by
INTRODUCTION 200 km (International Seismological Centre; see Data and Re-
sources; Fig. 1). The focal solutions of the USGS and Global
The relatively fast convergence between the oceanic Nazca Centroid Moment Tensor Project (Global CMT; Ekström
plate and South America in the region offshore of Ecuador et al., 2012) indicated that the earthquake involved primarily
and Colombia (e.g., García Cano et al., 2014) causes significant thrust faulting with minor strike-slip components on the in-
elastic strain to build up along the nonuniformly coupled plate terface of the plate boundary. The USGS finite-fault model
interface, resulting in several large megathrust earthquakes demonstrated a compact slip distribution overlapping, and ex-
(M w > 7) over the past century (Chlieh et al., 2014). One of tending slightly to the north, of the region that slipped during
the earliest megathrust earthquakes in this region was the great the 1942 M w 7.8 earthquake.

doi: 10.1785/0220160151 Seismological Research Letters Volume 88, Number 2A March/April 2017 1
SRL Early Edition

▴ Figure 1. Tectonic setting of the 2016 M w 7.8 Ecuador earthquake. Blue rectangles depict the footprints of the Sentinel-1A images used
in this study, and black lines depict active faults (Alvarado et al., 2014). Focal mechanism plots depict the focal mechanism solutions from
the Global Centroid Moment Tensor (Global CMT; Ekström et al., 2012) and U.S. Geological Survey (USGS). Yellow circles are two months
of aftershocks taken from the International Seismological Centre (ISC) catalog. Stars denote the epicenters of 1906–1998 M w > 7:0 earth-
quakes (Chlieh et al., 2014). Gray-shaded ellipses are the inferred high-slip region of the 1942, 1958, 1979, and 1998 earthquakes (Beck and
Ruff, 1984; Swenson and Beck, 1996; Segovia, 2001). Black dashed ellipse depicts the region of the megathrust inferred to have slipped in
the great 1906 earthquake, and dashed contours show the aftershocks zones of the 1942, 1958, and 1979 events (Mendoza and Dewey,
1984). The inset map shows the larger tectonic setting.

2 Seismological Research Letters Volume 88, Number 2A March/April 2017


SRL Early Edition

Table 1
Interferometry Pairs Used in This Study (A or D Denotes Ascending or Descending, Respectively)
Master Slave Inclined Azimuth
Track (A/D) (yyyy/mm/dd) (yyyy/mm/dd) Perp. B (m) Angle (°) Angle (°) σ (mm) α (km)
018 (A) 2016/03/29 2016/04/22 10 30–44 −168 8.2 15.8
040 (D) 2016/04/12 2016/04/24 18 30–44 −12 3.7 8.9
“Perp. B” is the perpendicular baseline of the pair; σ is the standard deviation calculated with all points in the nondeforming
area; and α is the e-folding correlation length scale of 1D covariance function (SAR data; see Data and Resources).

As demonstrated by numerous studies on large earth- as the 2015 M w 7.9 Gorkha, 2015 M w 8.3 Illapel, and 2015
quakes, spatially dense, near-field geodetic measurements of M w 6.4 Pishan earthquakes (Grandin et al., 2016; Solaro et al.,
coseismic surface deformation yields more insights into both 2016; Wen et al., 2016).
coseismic slip distribution and fault geometry than teleseismic We generate interferograms for both the ascending and
data (e.g., Lindsey et al., 2015; Zhang et al., 2015), although descending tracks (T018A and T040D) using the traditional
onshore geodetic data do have resolution issues when infer- two-pass differential InSAR method (e.g., Hanssen, 2001). All
ring offshore slip in megathrust earthquakes (e.g., Simons three subswaths within a satellite pass are combined to achieve
et al., 2011; Evans and Meade, 2012; Zhang et al., 2016). an efficient coverage of the coseismic deformation. We use the
Although there is a continuously operating Global Position- commercial processing software GAMMA (Werner et al.,
ing System (cGPS) network of 70 permanent stations in 2000), with geometric alignment based on precise orbit deter-
Ecuador (Mothes et al., 2013), the cGPS network is perhaps mination (SAR orbit data; see Data and Resources). Because
too sparse to characterize the full pattern of surface displace- straightforward interpolation of the slave image results in
ments associated with this earthquake. As a new generation banding with each burst, deramping the single look complex
satellite-based Synthetic Aperture Radar (SAR) system, the (SLC)-format images is done prior to interpolation to over-
European Space Agency’s Sentinel-1A has proven to be suc- come limitations of large phase mismatches (Xu et al., 2016).
cessful in determining ground deformation due to recent ma- We adopt the 90-m resolution Shuttle Radar Topography Mis-
jor earthquakes (e.g., Lindsey et al., 2015; Zhang et al., 2015; sion digital elevation model to remove topographic phases
Solaro et al., 2016). In this study, we process SAR images ac- (Farr et al., 2007). Finally, the interferograms are filtered using
quired by the new generation, C-band, Sentinel-1A SAR system a power spectrum filter algorithm (Goldstein and Werner,
and use these data to constrain both rupture geometry and slip 1998) to reduce the effects of phase noise, and they are un-
distribution of the 2016 earthquake. We also explore the reso- wrapped using the SNAPHU algorithm (Chen and Zebker,
lution of the Interferometric Synthetic Aperture Radar (InSAR) 2000). The unwrapped and geocoded interferograms have 90-m
data and discuss the earthquake within the context of the pre- resolution and represent range changes in the direction of line-
vious megathrust earthquakes offshore of Ecuador. of-sight (LOS) of the surface toward the satellite (Fig. 2a,b and
Ⓔ Fig. S1, available in the electronic supplement to this article).
INSAR DATA Because of short perpendicular baselines for the pairs of images,
there are no visible phase ramps in the resulting interferograms.
Onshore coseismic deformation in the 2016 Ecuador earth- The interferograms are shown in Figure 2, and they resolve
quake is imaged quite well by the Sentinel-1A satellite on both one incomplete lobe of coseismic deformation onshore. Even
ascending and descending orbits (Table 1). The Sentinel-1A though the coseismic deformation offshore is invisible by In-
radar operates in four modes with different resolution and cov- SAR, the onshore surface displacements are sufficient to con-
erage: strip-map mode (SM), interferometric wide swath mode strain the coseismic slip in this earthquake, as we show later.
(IW), extrawide swath mode (EW), and wave mode (WM). Compared to the ascending interferogram, the descending in-
Both IW and EW mode employ the novel terrain observation terferogram shows a clearer pattern of surface displacements
with progressive scans (TOPS) acquisition for interferometry because the descending image pairs are separated by a shorter
(Grandin et al., 2016), which achieves an extended swath by time span, resulting in a higher degree of coherence. The stan-
steering the beam in range, from backward to forward in the dard deviations of the LOS measurements in the region of the
azimuth direction for each burst (De Zan and Guarnieri, interferograms away from the prominent coseismic deforma-
2006). However, the TOPS success was realized at the cost tion (e.g., about 1° S and 1° N; Ⓔ Fig. S1) are 8.2 and
of a much higher accuracy of pixel co-registration at the level 3.7 mm for the ascending and descending interferograms, re-
of <0.001 pixel than is required to eliminate the azimuth phase spectively. We take these standard deviations as lower bounds
deramping (De Zan and Guarnieri, 2006). Currently, a tech- on the uncertainties of the interferometric data and set the
nique termed as enhanced spectral diversity performs well for a uncertainty of the LOS measurements to be 18 and 6 mm
single-frame data case (Xu et al., 2016). Sentinel-1A SAR data for the ascending and descending interferograms, respectively.
have been used for the investigation of recent earthquakes, such We note that range direction in the descending interferogram

Seismological Research Letters Volume 88, Number 2A March/April 2017 3


SRL Early Edition

▴ Figure 2. Coseismic interferograms of the (a) descending Sentinel-1A track T040D and the (b) ascending track T018A. Synthetic in-
terferograms on the (c) descending and (d) ascending tracks of the best-fit uniform slip-fault-plane solution; red barbed line represents
the surface trace of the modeled fault plane. Residual interferograms in (e) descending and (f) ascending geometries. Black line is the
subduction trench. Each contour on the interferograms indicates 0.05 m increment in displacement.

is approximately normal to the strike of the fault, which yields homogeneous elastic half-space on the inferred slip distribution
its maximum LOS offset of 73 cm, nearly double the maximum in the Discussion section.
offset (33 cm) in the ascending interferograms (Fig. 2a,b). We first determine the fault geometry through the
To reduce the computational burden when jointly inverting method of Clarke et al. (1997). We assume uniform slip on a
the two interferograms for coseismic slip, we subsample both rectangular fault plane, solving for fault strike, dip, length, and
interferograms using the QuadTree algorithm (Yan et al., 2013), width, as well as slip magnitude and rake. The method of
resulting in two data sets of 552 and 568 samples from the as- Clarke et al. (1997) uses a simplex search algorithm, starting
cending and descending interferograms, respectively. from an initial solution. We construct 100 initial models with
strike, dip, and rake randomly sampled from Gaussian proba-
COSEISMIC SLIP MODEL bility density functions (PDFs) with means corresponding to
the strike, dip, and rake of the nodal plane corresponding to
We determined the coseismic slip from the InSAR data in two the megathrust in the USGS focal mechanism and standard
steps. First, we use a nonlinear inversion to constrain the fault deviations of 15°. We randomly choose the fault length, width,
geometry, assuming uniform slip on a rectangular fault. Second, and slip in the initial models from Gaussian PDFs with means
we invert the InSAR data for the distribution of coseismic slip, given by dimensions consistent with the USGS magnitude
assuming either the fault plane determined in the first step or (80 km × 35 km × 2:5 m) and standard deviations of 1 m in
using a strike from the focal mechanism solutions of this earth- slip, 20 km in length, and 15 km in depth. After 50 restarts in
quake. We assume that the fault is embedded in a Poisson half- the simplex search, each initial model results in a unique sol-
space with shear modulus 32 GPa (Okada, 1992). We discuss ution, and we take the collection of final models obtained to
the resolution of these data and the ramifications of using a indicate the dependence on the initial starting models in the

4 Seismological Research Letters Volume 88, Number 2A March/April 2017


SRL Early Edition

Table 2
Source Parameters of the 2016 Ecuador Earthquake
Longitude Latitude Strike Dip Rake Depth Length Width Slip Moment Rms
Model (°) (°) (°) (°) (°) (km) (km) (km) (m) (×1020 N·m) M w (cm)
USGS-BW −79.926 0.352 26 16 113 20.6 – – – 7.054 7.8 –
Global −80.35 −0.16 26 23 123 24.1 – – – 5.530 7.8 –
CMT-SW
USM −80.97 0.03 8.7 17.1 110.7 – 80.8 28.3 2.54 6.32 7.83 5.8
±0.04 ±0.01 ±1.2 ±0.7 ±1.9 – ±0.5 ±1.6 ±0.06
Reference −80.97* 0.03* 8.7* 17.1* 106.7 – 180 140 2.48† 7.86 7.90 5.1
model
Preferred −80.97* 0.03* 26.0 17.1* 123.1 180 140 2.56† 7.15 7.87 4.1
model
Source Parameters of the 2016 Ecuador Earthquake Determined from the U.S. Geological Survey Body-Wave (USGS-BW) Focal
Mechanisms, the Global Centroid Moment Tensor (Global CMT) Surface-Wave Focal Mechanisms (Global CMT-SW), the Nonlinear
Inversion Assuming Uniform Slip Model (USM, Best-Fitting Parameters Are Given, along with the Standard Deviations of the Parameters
from the 100 Nonlinear Inversions Started with Unique Initial Models), a Linear Inversion Solving for the Distribution of Slip on the Best-
Fitting Fault Geometry in the Uniform Slip Model (Reference Model; Ⓔ Fig. S5), and a Linear Inversion Using the Focal Mechanism Strike
(Preferred Model; Fig. 3). Rms is root mean square between observation and simulation with model.
*The value is taken from the best-fitting USM model
†Maximum slip.

nonlinear inversion (Clarke et al., 1997). Most surprisingly, our grams. From this, we conclude that the InSAR alone, even
best-fit uniform slip model has a more north–south strike than with both ascending and descending orbits, cannot uniquely
either the USGS or Global CMT focal mechanisms (Table 2; resolve the strike of the fault. There is a trade-off between the
Ⓔ Fig. S1), and the 100 solutions from each chain of the non- strike and inferred slip rake in these two models, with mean
linear inversion all have strikes at least 10° more northerly from coseismic rake weighted by the slip magnitude of 123° and
the focal mechanism solutions (Ⓔ Fig. S2). 107° (90° corresponding to pure thrust motion of hanging
We next invert for the distribution of coseismic slip as- wall) using a strike of 26° or 8.7°, respectively, both of which
suming a planar fault geometry. As the best-fitting strike are consistent with the footwall moving about N83°E. That
found in the above nonlinear inversion is quite different from this footwall direction is coincident with the plate conver-
the strike of the trench, we consider two fault geometries. In gence and is independent of assumed fault strike indicates
both cases, the faults dip 17.1° to the east, which is the best that there is minimal, if any, slip partitioning of the subduc-
dip obtained in the nonlinear inversions above, and we as- tion convergence in this earthquake.
sume fault lengths and widths of 180 and 140 km, respec- We take the coseismic slip model with strike of 26° as our
tively, and divide the faults into 10 × 10 km fault patches preferred coseismic slip model (Figs. 3a and 4; we show the
(for a total of 252 patches). We separately consider two fault coseismic slip model and data fits using an 8.7° strike in Ⓔ
strikes, either a strike of 8.7°, the best fault strike obtained in Figs. S5 and S6). The preferred model contains a relatively
the nonlinear inversions above, or a strike of 26°, which is the compact region of oblique-thrust slip, with maximum slip of
strike of the nodal plane corresponding to the trench in the about 2.5 m at a depth of just under 20 km (corresponding to
USGS and Global CMT focal mechanism solutions (Table 2). about 65 km along the dip direction along the fault plane),
Because the primary mechanism of the earthquake was thrust with slip extending down to about 35 km depth (∼120 km
(revealed by both the moment tensor solution and the slip along dip). We infer a decrease in slip toward the trench, with
rake in our best-fit uniform slip model), we constrain the slip under 2 m of slip at the trench; however, the resolution near
rake within 85°–135°. We linearly invert for slip on each of the trench is the poorest with these onshore data, a point we
the two fault planes, assuming a second-order smoothing to address in the Discussion section. Our preferred coseismic slip
avoid unphysical oscillation of inferred slip, determining the model yields a geodetic moment of 7:15 × 1020 N·m, equiva-
penalty parameter on the regularization through a standard lent to M w 7.87. The moment of our preferred model is
L-curve analysis (Ⓔ Fig. S3; e.g., Aster et al., 2013). We find slightly higher than the USGS or Global CMT solutions,
that the solution using a strike of 26° is slightly better than although we note that the ascending and descending interfero-
using the strike found in the nonlinear inversions above (root grams contain 6 and 8 days, respectively, of the postseismic
mean square error of 4.1 cm vs. 5.1 cm), with no clear period and thus may contain a minor amount of ground
differences in the residuals of the downsampled interfero- deformation due to postseismic deformation.

Seismological Research Letters Volume 88, Number 2A March/April 2017 5


SRL Early Edition

▴ Figure 3. (a) Coseismic slip model determined from the interferograms in Figure 2, assuming a fault location and dip determined in the
nonlinear inversion and using the fault strike of 26° taken from the USGS and Global CMT focal mechanism solutions. Arrows show the
sense of motion of the hanging wall, and contours are labeled in meters. (b) Standard deviation of 100 coseismic slip models, each inferred
by adding random noise to the original interferograms prior to inversion.

DISCUSSION function of the interferograms, to the original interferograms,


and then invert those interferograms for the coseismic slip. We
Although the 2016 M w 7.8 Ecuador earthquake is the sixth then compute the standard deviation of the coseismic slip in
major megathrust earthquake (M w > 7) since the 1906 the 100 resulting slip models to quantify the uncertainty in
M w 8.8 Colombia–Ecuador earthquake, it is the largest mega- each subfault. We find that the standard deviation of slip is
thrust earthquake in Ecuador with coseismic slip to be mapped less than 2 cm over most of the fault but increases to just under
by InSAR data. Owing to the wide coverage (up to 250 km) 3 cm near the trench (Fig. 3b; Ⓔ Fig. S5b). In a synthetic
and short satellite repeat time (12 days), the TOPS mode of the sensitivity test (sensitivity tests are described in the Ⓔ elec-
Sentinel-1A satellite provides an excellent image of the coseis- tronic supplement), we find that when coseismic slip is
mic deformation in both the descending and ascending orbits. constrained to end 40 km from the trench along the mega-
In comparison with the 70 GPS continuous stations in Ecua- thrust, we infer about 1 m of coseismic slip near the trench
dor (Mothes et al., 2013), the InSAR measurements are using onshore geodetic data (Ⓔ Fig. S7). The inferred slip in
spatially dense and fully cover the coseismic deformation of this sensitivity test is about half the magnitude of largest near-
this large earthquake with high precision. In addition, the orbit trench slip in our preferred model, suggesting that not all of the
accuracy of Sentinel-1A has been improved in contrast to near-trench slip in our preferred model is due to lack of data
advanced synthetic aperture radar (ASAR) data, so the orbital resolution. Using a layered elastic structure instead of the
phase ramp is insignificant and can be easily removed using a homogeneous elastic half-space we assumed in our preferred
constant offset or simple linear model (Feng et al., 2016). slip model (Fig. 3), we infer less slip near the trench (just over
10 cm lower magnitude slip) and slightly more slip extending
to the north (Ⓔ Fig. S8; details of the elastic structure are
Resolution of Near-Trench Coseismic Slip
presented in the Ⓔ electronic supplement). Based on the three
The 2016 earthquake generated a minor tsunami (e.g., Ye et al.,
points discussed here, we conclude that our coseismic model is
2016). We do resolve about 1–2 m of coseismic slip near the
consistent with slip extending to the trench, although we can-
trench, although it is important to note that with these
not uniquely constrain the magnitude of near-trench slip with
onshore data our resolution is the poorest near the trench. To
these data alone.
illustrate the degree to which these data are able to constrain
the coseismic slip near the trench, (1) we assess the error in our
slip model using a Monte Carlo-based resampling technique The 2016 Earthquake in Context of Past Seismicity
(e.g., Parsons et al., 2006), (2) we invert synthetic slip models Our inferred coseismic slip for the 2016 earthquake coincides
with and without near-trench coseismic slip, and (3) we rein- with the region of the megathrust that slipped in the
vert for coseismic slip using a layered elastic structure. To 1942 M 7.8 earthquake (Fig. 5). Swenson and Beck (1996)
estimate the uncertainty in our inferred slip, we add 100 real- inferred that the 1942 earthquake ruptured a region about
izations of correlated noise, simulated using the 1D covariance 50 km × 50 km, based on their source time function of

6 Seismological Research Letters Volume 88, Number 2A March/April 2017


SRL Early Edition

▴ Figure 4. Modeled interferograms for the (a) descending track T040D and (b) ascending track T018A predicted by the slip model in
Figure 3a and the associated residuals of the observed interferograms in Figure 2c,d. The red line is the trace of the model fault plane, and
the white dashed lines represent the surface projections of the modeled fault. Each contour indicates a 0.05-m increment in displacement.
These interferograms are also shown wrapped in Ⓔ Figure S4 (available in the electronic supplement to this article).

the earthquake and an assumed rupture velocity of 2:0 km=s. would be required to have ruptured to match the moment. A
Using this rupture size, however, the average slip would need 50-km circle approximating the 1942 earthquake roughly cor-
to be greater than 8 m to match the moment (Swenson and responds to the region in which we infer more than 1.0 m of
Beck, 1996). If on the other hand, we assume uniform slip of slip in our preferred model (Fig. 5). The largest MMI (IX)
2.0 m, consistent with the slip we infer for this earthquake, in the 1942 earthquake was reported to be just to the south
then a region of the megathrust with radius about 50 km of the epicenter, with MMI ≥ VIII over the coastal region

Seismological Research Letters Volume 88, Number 2A March/April 2017 7


SRL Early Edition
cially given the differences between reported MMI in this and
the 1942 events, that details of the coseismic slip in each of the
events were not exactly the same, as also concluded by Ye et al.
(2016). Even if there are differences in the exact pattern of slip
in the 1942 and 2016 earthquakes, it is reasonable to assume
that these two earthquakes involved slip within the same asper-
ity given the overlap in slip in the two earthquakes and the fact
that on average the slip in these two earthquakes are consistent
with 110 years of slip deficit.
White et al. (2003) argued that the megathrust offshore
Ecuador was only ∼50% locked, based on onshore GPS data. It
is worth noting that although White et al. (2003) concluded
that only part of the convergence was accommodated by slip in
large megathrust earthquakes, they only considered a slip mag-
nitude of the 1942 earthquake determined by Swenson and
Beck (1996) by assuming uniform slip over the entire after-
shock zone. However, Swenson and Beck (1996) argued that
the aftershock zone is likely much larger than the region that
slipped in 1942 and thus that slip was likely higher than 1.2 m
if the entire aftershock region coseismically slipped. In the
2016 earthquake, aftershocks also extended to the north of the
▴ Figure 5. Preferred coseismic slip model of the 2016 Ecuador coseismic slip, well past Esmeraldas, Ecuador (Fig. 5). Chlieh
earthquake (contours) and Global CMT centroid location (gold
et al. (2014) found that the region corresponding to the 1942
star), along with approximate region of the megathrust that
earthquake was also only partially coupled, although there was
slipped in the 1942 earthquake (Swenson and Beck, 1996), rep-
a fair amount of variability of the exact coupling depending on
resented as either 25 or 50 km radius circles with either ∼9 or
the regularization. The partially locked view of this region in
2.2 m of uniform slip, respectively (black circles). Blue dots
Chlieh et al. (2014) is seemingly contradictory to our above ar-
are the relocated aftershocks of the 1942, 1958, and 1979 earth-
gument of a single locked asperity that ruptured in the 1942 and
quakes (Mendoza and Dewey, 1984); gold dots are two months of
aftershocks of the 2016 earthquake from ISC (see Data and Re-
2016 earthquakes, which also likely ruptured in the great 1906
sources); black line is the trench; yellow squares are prominent
earthquake. However, it is important to note that this region
Ecuadorian cities; black dashed line is the modified Mercalli in-
suffers from a relative paucity of data, with only three GPS sta-
tensity (MMI) VIII contour of the 1942 earthquake (Swenson and
tions farther from the coast around the equator. The poor res-
Beck, 1996); and reported MMI in the 2016 earthquake are indi- olution of the GPS data is perhaps manifested in the large range
cated in black roman numerals (USGS). of inferred coupling that Chlieh et al. (2014) find in the vicinity
of the 1942 epicenter among the various regularizations they
considered (compare to fig. 5 of Chlieh et al., 2014).
(Swenson and Beck, 1996; Fig. 5). MMI reported in the 2016 It is important to note that we are unlikely to resolve
earthquake was lower (maximum MMI of VIII to the east subtle variations in coseismic slip using these InSAR data and
and slightly north of the epicenter; USGS, see Data and Re- the level of smoothing chosen in our inversions. The USGS
sources), and falling to IV at Quito (Fig. 5), likely indicating finite-slip model and the model of Ye et al. (2016), both of
that the details of the slip distribution in these two earth- which rely on seismological data, infer more heterogeneous
quakes differed. coseismic slip in the 2016 earthquake than we infer. In the
Assuming that the 2016 and 1942 events represent failure extreme case in which the coseismic slip is constrained in two
within the same asperity, which also ruptured during the 1906 small patches separated by 20 km, using synthetic interfero-
earthquake, an interesting question is whether the cumulative grams and the smoothing constraints in our preferred model,
amount of slip in these two events matches the accumulated we cannot resolve these two distinct slip patches in a sensitivity
slip deficit since 1906. With a convergence rate of 56 mm=yr test (Ⓔ Fig. S4), suggesting that subtle variations in the slip as
(e.g., García Cano et al., 2014), there has been just over 6 m of inferred by Ye et al. (2016) would not be resolved by only the
slip deficit accumulated on the megathrust over the 110 years onshore interferograms.
since the great 1906 earthquake. An earthquake with uniform The region located between the 1942 and 1958 events has
slip of about 3 m over a ∼40km radius circular patch is equiv- been devoid of major seismicity, and this seismic gap has been
alent to a moment magnitude 7.8 (∼50 km radius for M 7.9). suggested to have potential for future ruptures (e.g., Papadimi-
Although this slip is larger than the maximum slip we infer in triou, 1993; Beauval et al., 2013; Chlieh et al., 2014; García
the 2016 earthquake, it is important to note that the imposed Cano et al., 2014). As the recent 2016 earthquake also did
smoothing regularization likely results in underestimated slip not rupture this seismic gap, there may be an ongoing earth-
over a broader region of the fault. It may also be likely, espe- quake risk within the seismic gap. It is interesting to note that

8 Seismological Research Letters Volume 88, Number 2A March/April 2017


SRL Early Edition
aftershocks of both the 1942 and 2016 earthquakes extend over https://www.unavco.org/data/imaging/sar/lts1/winsar/s1qc/
this seismic gap (Mendoza and Dewey, 1984; Swenson and aux_poeorb/ (last accessed August 2016).
Beck, 1996; Fig. 5), and Chlieh et al. (2014) observed the meg-
athrust in this region being at least partially coupled. Ongoing
geodetic observations, particularly InSAR measurements with ACKNOWLEDGMENTS
higher spatial coverage than GPS, can test for the presence of
afterslip within the seismic gap following the 2016 earthquake All figures were generated using Generic Mapping Tools
and thereby help resolve the degree to which strain is accumu- (GMT). This work is supported by National Natural Science
lating in the region to the north of the 1942/2016 earthquake. Foundation of China (41204001, 41274037, and 41431069),
China Postdoctoral Science Foundation (2015M57228), the
Basic Fund of Hubei Subsurface Multi-scale Imaging Key Lab-
CONCLUSION oratory, Institute of Geophysics and Geomatics, China Univer-
sity of Geosciences,Wuhan (SMIL-2015-01), the Fundamental
We obtain interferograms showing ground deformation during Research Funds for National Universities (CUGL150810),
the 2016 M w 7.8 Ecuador earthquake, along descending and and the Basic Research Fund of Key Laboratory of Geospace
ascending Sentinel-1 SAR orbits. We invert these interfero- Environment and Geodesy, Ministry of Education of China
grams to infer an image of coseismic slip in this earthquake. We (13-02-11 and 14-01-01). Ping He is a visiting scientist at
find oblique thrust slip, with maximum slip about 2.5 m. Co- the University of Michigan, sponsored by the State Scholarship
seismic slip over 2 m is in the depth range of about 10–25 km, Fund of China.
with the maximum slip at about 20-km depth. Our results
suggest that there is little, if any, partitioning of the oblique
plate convergence. The coseismic slip in the 2016 earthquake REFERENCES
is coincident with the region inferred by Swenson and Beck
(1996) to have slipped in the M w 7.8 1942 earthquake. The Alvarado, A., L. Audin, J. M. Nocquet, S. Lagreulet, M. Segovia, Y. Font,
reported MMI in the two events differ, suggesting that and P. Jarrin (2014). Active tectonics in Quito, Ecuador, assessed by
although the 2016 earthquake involves the failure of the same geomorphological studies, GPS data, and crustal seismicity, Tectonics
33, no. 2, 67–83.
asperity that ruptured in the 1942 earthquake, there may be Aster, R. C., B. Borchers, and C. H. Thurber (2013). Parameter Estima-
some differences in the details of slip in the two earthquakes. tion and Inverse Problems, Second Ed., Academic Press, Socorro,
Assuming the 1906, 1942, and 2016 events, all ruptured the New Mexico, 376 pp.
same asperity (with the larger 1906 earthquake also involving Beauval, C., H. Yepes, P. Palacios, M. Segovia, A. Alvarado, Y. Font, and S.
slip on multiple asperities), the average repeat time for failure Vaca (2013). An earthquake catalog for seismic hazard assessment in
Ecuador, Bull. Seismol. Soc. Am. 103, no. 2A, 773–786.
of this asperity is 55 years. This estimate should be taken with Beck, S. L., and L. J. Ruff (1984). The rupture process of the great 1979
caution, because it is simply the mean of two repeat times. A Colombia earthquake: Evidence for the asperity model, J. Geophys.
recurrence time of 55 yrs is much shorter than 140  30 yrs Res. 89, no. B11, 9281–9291.
suggested by Chlieh et al. (2014), although is in line with Chen, C. W., and H. A. Zebker (2000). Network approaches to
repeat times suggested by Nishenko (1991) for nearby seg- two-dimensional phase unwrapping: Intractability and two new
algorithms, J. Opt. Soc. Am. 17, 401–414.
ments for which more are known. If repeat times of other Chlieh, M., P. A. Mothes, J. M. Nocquet, P. Jarrin, P. Charvis, D. Cis-
nearby asperities are similar to this, the Esmeraldas region neros, and M. Vallée (2014). Distribution of discrete seismic
near the 1958 earthquake is also a likely candidate for a future asperities and aseismic slip along the Ecuadorian megathrust, Earth
earthquake. Planet. Sci. Lett. 400, 292–301.
Clarke, P. J., D. Paradissis, P. Briole, P. C. England, B. E. Parsons, H.
Billiris, and J. C. Ruegg (1997). Geodetic investigation of the
DATA AND RESOURCES 13 May 1995 Kozani-Grevena (Greece) earthquake, Geophys.
Res. Lett. 24, no. 6, 707–710.
The Sentinel-1A SAR data were downloaded from the Senti- De Zan, F., and A. M. Guarnieri (2006). TOPSAR: Terrain observation
by progressive scans, IEEE Trans. Geosci. Remote Sens. 44, no. 9,
nel-1A Scientific Data Hub (https://vertex.daac.asf.alaska. 2352–2360.
edu/, last accessed June 2016). The moment tensor solution is Ekström, G., M. Nettles, and A. M. Dziewoński (2012). The global CMT
from the U.S. Geological Survey (USGS; http://earthquake. project 2004–2010: Centroid-moment tensors for 13,017 earth-
usgs.gov, last accessed August 2016) and Global Centroid Mo- quakes, Phys. Earth Planet. In. 200, 1–9
ment Tensor Project (Global CMT; http://www.globalcmt.org , Evans, E. L., and B. J. Meade (2012). Geodetic imaging of coseismic slip
and postseismic afterslip: Sparsity promoting methods applied to
last accessed August 2016; Fig. 1). The aftershock locations are the great Tohoku earthquake, Geophys. Res. Lett. 39, L11314,
from International Seismological Centre (ISC; http://www.isc. doi: 10.1029/2012GL051990.
ac.uk, last accessed August 2016). The 2016 Ecuador earth- Farr, T. G., P. A. Rosen, E. Caro, R. Crippen, R. Duren, S. Hensley, and
quake report is from http://earthquake-report.com (last ac- D. Seal (2007). The shuttle radar topography mission, Rev. Geophys.
cessed June 2016). The details on the International Disaster 45, no. RG2004, doi: 10.1029/2005RG000183.
Feng, G., Z. Li, B. Xu, X. Shan, L. Zhang, and J. Zhu (2016). Coseismic
Database (EM-DAT), Centre for Research on the Epidemiol- deformation of the 2015 M w 6.4 Pishan, China, earthquake
ogy of Disaster, can be found at http://www.emdat.be/ (last estimated from Sentinel-1A and ALOS2 data, Seismol. Res. Lett.
accessed October 2016). SAR orbit data can be found at 87, no. 4, 1–7.

Seismological Research Letters Volume 88, Number 2A March/April 2017 9


SRL Early Edition
García Cano, L. C., A. Galve, P. Charvis, and B. Marcaillou (2014). White, S. M., R. Trenkamp, and J. N. Kellogg (2003). Recent crustal
Three-dimensional velocity structure of the outer fore arc of the deformation and the earthquake cycle along the Ecuador-Colombia
Colombia–Ecuador subduction zone and implications for the subduction zone, Earth Planet. Sci. Lett. 216, no. 3, 231–242.
1958 megathrust earthquake rupture zone, J. Geophys. Res. 119, Xu, B., Z. Li, G. Feng, Z. Zhang, Q. Wang, J. Hu, and X. Chen (2016).
no. 2, 1041–1060. Continent-wide 2-D co-seismic deformation of the 2015 M w 8.3
Goldstein, R. M., and C. L. Werner (1998). Radar interferogram filter- Illapel, Chile earthquake derived from Sentinel-1A data: Correction
ing for geophysical applications, Geophys. Res. Lett. 25, no. 21, of azimuth co-registration error, Remote Sens. 8, no. 5, 376.
4035–4038. Yan, Y., V. Pinel, E. Trouvé, E. Pathier, J. Perrin, P. Bascou, and F. Jouanne
Grandin, R., E. Klein, M. Métois, and C. Vigny (2016). Three-dimensional (2013). Coseismic displacement field and slip distribution of
displacement field of the 2015 M w 8.3 Illapel earthquake (Chile) the 2005 Kashmir earthquake from SAR amplitude image
from across- and along-track Sentinel-1 TOPS interferometry, Geo- correlation and differential interferometry, Geophys. J. Int. 193,
phys. Res. Lett. 43, no. 6, 2552–2561. no. 1, 29–46.
Hanssen, R. (2001). Radar Interferometry: Data Interpretation and Error Ye, L., H. Kanamori, J. P. Avouac, L. Li, K. F. Cheung, and T. Lay (2016).
Analysis, Kluwer Academic Publishers, Amsterdam, The Nether- The 16 April 2016, M w 7.8 (M s 7.5) Ecuador earthquake: A quasi-
lands. repeat of the 1942 M s 7.5 earthquake and partial re-rupture of the
Lindsey, E. O., R. Natsuaki, X. Xu, M. Shimada, M. Hashimoto, D. 1906 M s 8.6 Colombia–Ecuador earthquake, Earth Planet. Sci.
Melgar, and D. T. Sandwell (2015). Line-of-sight displacement from Lett. 454, 248–258.
ALOS-2 interferometry: M w 7.8 Gorkha Earthquake and M w 7.3 Zhang, G., E. Hetland, and X. Shan (2015). Slip in the 2015 M w 7.9
aftershock, Geophys. Res. Lett. 42, no. 16, 6655–6661. Gorkha and M w 7.3 Kodari, Nepal, earthquakes revealed by seismic
Mendoza, C., and J. W. Dewey (1984). Seismicity associated with the and geodetic data: Delayed slip in the Gorkha and slip deficit
great Colombia–Ecuador earthquakes of 1942, 1958, and 1979: between the two earthquakes, Seismol. Res. Lett. 86, no. 6,
Implications for barrier models of earthquake rupture, Bull. Seismol. 1578–1586.
Soc. Am. 74, no. 2, 577–593. Zhang, Y., G. Zhang, E. A. Hetland, X. Shan, S. Wen, and R. Zuo (2016).
Mothes, P. A., J. M. Nocquet, and P. Jarrín (2013). Continuous GPS Coseismic fault slip of the September 16, 2015 M w 8.3 Illapel, Chile
network operating throughout Ecuador, Eos Trans. AGU 94, earthquake estimated from InSAR data, Pure Appl. Geophys. 173,
no. 26, 229–231. 1029, doi: 10.1007/s00024-016-1266-3.
Nishenko, S. P. (1991). Circum-Pacific seismic potential: 1989–1999,
Pure Appl. Geophys. 135, 169–259. Ping He1
Okada, Y. (1992). Internal deformation due to shear and tensile faults in a
half-space, Bull. Seismol. Soc. Am. 82, no. 2, 1018–1040. Qi Wang
Papadimitriou, E. E. (1993). Long-term earthquake prediction along the Rong Zou
western coast of South and Central America based on a time Hubei Subsurface Multi-scale Imaging Key Laboratory
predictable model, Pure Appl. Geophys. 140, no. 2, 301–316. Institute of Geophysics and Geomatics
Parsons, B., T. Wright, P. Rowe, J. Andrews, J. Jackson, R. Walker, and E. China University of Geosciences
R. Engdahl (2006). The 1994 Sefidabeh (eastern Iran) earthquakes
revisited: New evidence from satellite radar interferometry and car- Wuhan 430074, China
bonate dating about the growth of an active fold above a blind wangqi@cug.edu.cn
thrust fault, Geophys. J. Int. 164, no. 1, 202–217, doi: 10.1111/
j.1365-246X.2005.02655.x. Eric A. Hetland
Segovia, M. (2001). El sismo de Bahía del 4 de agosto de 1998: Carac- Department of Earth and Environmental Sciences
terización del mecanismo de ruptura y análisis de la sismicidad en la
zonacostera, Titulo de Ingeniera Geologiatesis, Escuela Politecnica University of Michigan
Nacional, Quito, Ecuador, 126 pp. (in Spanish). Ann Arbor, Michigan 48109 U.S.A.
Simons, M., S. E. Minson, A. Sladen, F. Ortega, J. Jiang, S. E. Owen, L. ehetland@umich.edu
Meng, J.-P. Ampuero, S. Wei, R. Chu, et al. (2011). The 2011
Magnitude 9.0 Tohoku-Oki earthquake: Mosaicking the megathrust Kaihua Ding
from seconds to centuries, Science 322, 1421, doi: 10.1126/sci-
ence.1206731. Faculty of Information Engineering
Solaro, G., V. De Novellis, R. Castaldo, C. De Luca, R. Lanari, M. Man- China University of Geosciences
unta, and F. Casu (2016). Coseismic fault model of M w 8.3 2015 Wuhan 430074, China
Illapel earthquake (Chile) retrieved from multi-orbit Sentinel1-A
DInSAR measurements, Remote Sens. 8, no. 4, 323. Yangmao Wen
Swenson, J. L., and S. L. Beck (1996). Historical 1942 Ecuador and 1942
Peru subduction earthquakes and earthquake cycles along Colom- School of Geodesy and Geomatics
bia–Ecuador and Peru subduction segments, Pure Appl. Geophys. Wuhan University
146, no. 1, 67–101. Wuhan, Hubei 430079, China
Wen, Y., C. Xu, Y. Liu, and G. Jiang (2016). Deformation and source
parameters of the 2015 M w 6.5 earthquake in Pishan, western China, Published Online 25 January 2017
from Sentinel-1a and ALOS-2 data, Remote Sens. 8, no. 2, 134.
Werner, C., U. Wegmüller, T. Strozzi, and A. Wiesmann (2000). Gamma 1
SAR and interferometric processing software, Proc. of the ERS-Envi- Department of Earth and Environmental Sciences, University of Mich-
sat Symposium, Gothenburg, Sweden, 16 October 2000, 1620 pp. igan, Ann Arbor, Michigan 48109 U.S.A.

10 Seismological Research Letters Volume 88, Number 2A March/April 2017

You might also like