You are on page 1of 9

J. Earth Syst. Sci.

(2021)130:44 Ó Indian Academy of Sciences


https://doi.org/10.1007/s12040-020-01531-9 (0123456789().,-volV)(0123456789(
).,-volV)

A reliable velocity estimation in a complex deep-water


environment using downward continued long oAset
multi-channel seismic (MCS) data

DIBAKAR GHOSAL1,* and SATISH C SINGH2


1
Indian Institute of Technology Kanpur, Kanpur 208 016, UP, India.
2
Institut de Physique du Globe de Paris, Jussieu, 75014 Paris, France.
*Corresponding author. e-mail: dghosal@iitk.ac.in
MS received 14 May 2020; revised 11 August 2020; accepted 28 October 2020

The estimation of a reliable velocity–depth model from towed streamer marine seismic data recorded in deep
water, especially with a complex seaCoor environment, is challenging. The determination of interval velocities
from the normal move-out (NMO) of the reCected seismic signals for shallow reCectors (\1 km below the
seaCoor) is compromised by the combination of a long wave path in the water column and the complex ray
paths due to topography, leading to small move-out differences between reCectors. Furthermore, low sediment
velocities and deep water produce refraction arrivals only at limited far oAsets that contain information about
deeper structures. Here, we present an innovative method where towed streamer seismic data are downward
continued to the seaCoor leading to the collapse of the seaCoor reCection and the emergence of refraction events
as Brst arrivals close to zero oAset, which are used to determine a high-resolution near surface velocity–depth
model using an eDcient tomographic method. These velocities are then used to perform pre-stack
depth migration. We found that the velocity–depth model derived from tomography of downward continued
towed streamer data provides a far superior pre-stack depth migrated image than those produced from
velocity–depth models derived from conventional velocity estimation techniques.
Keywords. Downward continuation; tomography; pre-stack depth migration; OBS; MCS.

1. Introduction rocks that do not produce any coherent reCections


and thus distort deeper reCections which make it
Conventional velocity–depth model building using difBcult to determine the near surface velocity. To
normal moveout (NMO) based velocity analysis overcome such difBculties, we present here an appli-
technique has several limitations in deep water cation of the approach described in Ghosal et al.
environment like in the Gulf of Mexico (Galloway (2012, 2014), in which downward continuation of
2008) where imaging of salt diapirs is a real challenge; surface seismic data to the seaCoor is done, followed
and in subduction settings such as Cascadia (Yelisetti by a travel time tomography of Brst arrivals.
et al. 2014), Sumatra (Singh et al. 2012) where the We demonstrate the applicability of the method
seaCoor could be very rough and water depth could on a long oAset dataset from the northern oAshore
vary drastically from a couple of meters to several Sumatra where the Indian plate is obliquely sub-
kilometers across the subduction front. Moreover, the ducting at a rate of 52 mm/yr (Bgure 1) beneath the
top parts of the shallow sub-seaCoor may consist of Eurasian plate (Fitch 1972; Liu et al. 1983; Matson
clastics, salt diapirs or highly deformed sedimentary and Moore 1992; Sieh and Natawidjaja 2000;
44 Page 2 of 9 J. Earth Syst. Sci. (2021)130:44

B
A

Figure 1. Location of seismic reCection proBle WG2 superimposed on recently acquired bathymetric data. The speciBc seismic
data used for the migrated images was from the whitened segment of WG2 as the line was shot over the eastern rise onto the
accretionary. A, B, C are positions of shot gathers from 5.5 to 12 km streamers shown in Bgure 2.

McCaffrey 2009). This region, therefore, is associ- 2. Datasets


ated with several major fault systems including
Great Sumatran fault, West Andaman fault and In July–August 2006, coincident deep seismic
Mentawai fault and have produced several dev- refraction and reCection proBles were acquired
astating major to great tsunamigenic earthquakes using the French R/V Marion Dufresne and the
such as 26th Dec, 2004 ‘boxing day’ earthquake WesternGeco M/V Geco Searcher vessel carrying
with magnitude 9.3. Despite of geological com- 8260 cubic inch and 10170 cubic inch airgun array
plexity, these data hold extreme significance sources, respectively. The 520-km long WG2 pro-
owing to 3–5 km thick sediment-Blled trench at Ble (Bgure 1) oriented *20° anticlockwise from the
4.5 km water depth and, a striking deformation trench normal on which 56 OBSs spaced at 8.1 km
front and accretionary prism (Bgure 1) rising to 2 were deployed and shots were Bred at 150 m
km water with many deep-rooted faults (Ghosal intervals. A 12-km-long streamer towed at 15 m
et al. 2014). depth was used for the reCection survey at 50 m
Primarily velocity–depth models are estimated shot interval, to image the deeper structure of the
using three different methods: the NMO based con- megathrust (Singh et al. 2008, 2012; Ghosal et al.
ventional processing technique, travel time inversion 2012, 2014). Another 5.5-km-long streamer was
of ocean bottom seismometer (OBS) data, and towed at 7.5 m water depth to obtain high-resolu-
downward continued multi-channel seismic (MCS) tion data to better image the near surface features.
streamer data. The accuracy of velocity models Since the depth of the short streamer is less, the
is demonstrated by performing pre-stack depth recorded ghost energy produced a notch at 100 Hz
migration. and thus a useful bandwidth of 20–80 Hz is easily
J. Earth Syst. Sci. (2021)130:44 Page 3 of 9 44

(a) Distance (km) (b) Distance (km) In this study, the 5.5 km towed streamer data
0 3 5.5 0 3 6 9 12 are used to produce the pre-stack depth migrated
Swell noise
images; the three velocity–depth models are
6 derived from the 5.5 km towed streamer data, the
Refracted
Time (s)

OBS data and the longer 12.0 km towed streamer


8 data.
Refraction

10 3. Methodology
(c) (d)
5 3.1 Conventional processing of 5.5 km
streamer data
Time (s)

7 Refraction The processing of short streamer data started


removing low frequency swell noise (0–20 Hz) using
9 f-k Bltering (SheriA and Geldart 1982). The sam-
pling interval of the short streamer data is kept
(e) (f) intact as 2 ms. After the pre-processing, the data-
5 sets are sorted into common mid-point (CMP)
Time (s)

Refraction gather and supergather is formulated using eight


7 adjacent CMPs (Bgure 3a, b) in which the sem-
Multiples blance analysis (Bgure 3d) is carried out to obtain
9 Multiples the root mean square (RMS) velocity–depth func-
tion, which is further converted into interval
Figure 2. Raw shot gathers from 5.5 km (a, c and e) and 12 velocity (Bgure 3d). The RMS velocity–depth
km (b, d and f) long streamers at locations A (a, b), B (c, d) function is further used for normal move-out
and C (e, f) shown in Bgures 1 and 6. (NMO) and stacking (Bgure 3c) of the CMP
gathers. The stack section is, further, migrated
used for providing high-resolution seismic image of using KirchhoA time migration algorithm to obtain
the near surface. On the other hand, for the long the image of the subsurface in which an AGC gain
deeper streamer data, the ghost notch appears in with 1000 ms window length is applied to highlight
the frequency spectrum at 50 Hz and thereby the interesting subsurface details.
low frequency energy essentially enclosed between
8 and 40 Hz is useful for deep seismic imaging
(Singh et al. 2012). 3.2 Advanced processing of MCS data
Figure 2 shows a comparison between the
acquired long and short streamer datasets across The advanced processing is carried out in three
the subduction front using three shot gathers A, B steps. At Brst, we have carried out the downward
and C (Bgure 1) located at 5 km (Bgure 2a and b), continuation of the long streamer data followed by
12 km (Bgures 2c and d) and 23 km (Bgures 2e and travel time tomography of these data and pre-stack
f) distances, respectively. The position of these depth migration using the tomography derived
shot gathers (A, B, C) are marked in Bgures 1 and velocity–depth model. The details of these tech-
6. Both the streamer datasets were aAected by the niques are explained in the following sections:
low frequency swell noise and water bottom mul-
tiples (Bgure 2e–f). The recorded reCected phases 3.2.1 Downward continuation
show clear hyperbolic trend at the trench for both
the streamer datasets, but become more compli- Towed marine seismic data recorded over deep
cated along the updip direction (Bgure 2e–f). The water often leads to a superposition of reCection and
traces of refracted phases at the short streamer refraction events from near-surface structure on the
data are completely missing in the study area, recorded shot records. Downward continuation
whereas these arrivals are recorded in the long (DC) (Berryhill 1979; Arnulf et al. 2011; Ghosal et al.
streamer at *10.5–12 km (Bgure 2b) and *9–12 2012, 2014) extrapolates the water-surface recorded
km oAset (Bgure 2f) across the front with decreas- waveBeld to an arbitrary surface, e.g., seaCoor, so
ing water depth. that refraction events become the Brst arrivals
44 Page 4 of 9 J. Earth Syst. Sci. (2021)130:44

(a) Supergather trace no. (b) Supergather trace no. (c) Stack (d)
14 134 254 374 14 134 254 374 1 2 3 4 5
5
5 5 5

Interval
velocity
6 6 6
6

Time (s)

Time (s)
Time (s)

Time (s)

7 7 7 7

8 8 8 8

Figure 3. (a) Processed CMP supergather at location B from the subduction front, (b) NMO corrected supergather, (c) local
stack sections, (d) velocity analysis using semblance method, the interval velocity is marked in black.

Figure 4. A schematic diagram explains the eAect of downward continuation of towed streamer seismic data so that the source
and receiver positions are located on the seaCoor.

across the shot record’s oAset range and reCections and (3) getting solutions to wave equation using
arrive after their associated refractions (Bgure 4). Bnite difference method (Claerbout 1976).
The seaCoor reCection is also removed from the In the present study, we have followed the
downward continued seismic shot record, leading a second approach, i.e., the KirchhoA’s integral
survey geometry similar to land seismic survey method in which the recorded waveBeld at the
without any ground roll (Arnulf et al. 2011; Ghosal water surface is convolved with the time lagged
et al. 2012, 2014). Apart from bringing the refracted delta function and the result is summed over all the
arrival as Brst arrivals, the DC method collapses shot gathers (Harding et al. 2007). The time delays
seaCoor diAractions, thus, improving the imaging are calculated dividing the distance between the
condition (Arnulf et al. 2011). Three different ways surface positions of shots and hydrophones and
are practically used to carry out downward contin- their extrapolated positions on the seaCoor by the
uation of the recorded waveBelds that are: (1) using constant water velocity (i.e., 1480 m/s). The entire
pre-stack phase-shift method in the frequency DC process is carried out in three steps: (1) the
domain (Gazdag 1978), (2) applying KirchhoA’s receiver locations for different shots are extrapo-
integral method in the time domain (Berryhill 1979), lated vertically down to the seaCoor, (2) the
J. Earth Syst. Sci. (2021)130:44 Page 5 of 9 44

extrapolated shot gathers are sorted into common are extrapolated to the seaCoor. The refracted pha-
receiver location gathers (CRL) and full folded ses, which are observed at very far oAset between 9
gathers are separated out, and (3) then the shot and 12 km (Bgure 2d), are now visible at the near
locations are extrapolated to the seaCoor using the oAset extending until the end of the streamer. The
full folded CRL gathers. Once both the shots and direct arrivals and the eAect of the seaCoor reCec-
receivers are extrapolated to seaCoor, we sort back tions are thus superseded. The downward continued
the extrapolated full folded CRL gathers into shot shot records are, therefore, easier to interpret and
gathers. If the shot intervals are more than 25 m, make velocity–depth model building through Brst
the aliasing eAect may dominate the CRL gathers arrival tomography more accurate.
and thus should be handled by interpolating the
shot intervals to either 25 or 12.5 m after the step 2.
The improvement gained in reCection and 3.2.2 Travel time tomography of downward
refraction event discrimination by DC of the source continued 12-km long data
and receiver positions to the seaCoor is illustrated in
Bgure 5. Figure 5(a) shows the impact of downward Travel time tomography, which provides a smooth
continuation on the 12.0-km long streamer shot velocity–depth model, was carried out by inverting
record located on the frontal slope at location B refracted arrivals from every fourth downward
(Bgure 2d) in which the source and receiver positions continued shot gathers using the algorithm devel-
oped by Van Avendonk et al. (2004). The Brst
arrivals from those shot gathers are picked
(Bgure 5a) using a semi-automatic technique,
maintaining consistency between two consecutive
shot gathers, keeping the preceding travel time
function as a guided function. The uncertainty is
calculated based on several factors, for instance, (1)
positions of shots and receivers on the water surface
contribute *2 ms, (2) bathymetry eAect con-
tributes *10 ms, (3) extrapolated positions of shots
and receivers cause *8 ms, and (4) picking uncer-
tainties include *8 ms, and thus the total uncer-
tainty calculated as 15 ms. The forward modelling
is conducted using shortest path method (Moser
1991) where the grid sizes are kept as 50 m and the
starting model is used as the pre-existing 2D
velocity–depth model obtained from travel time
tomography of OBS data (Chauhan et al. 2009;
Chauhan 2010; Singh et al. 2012). During inversion,
the misBt between the observed and predicted is
minimized using a least square approach, regularizing
the inversion scheme by a roughness matrix, con-
sisting of Brst and second order derivatives, which are
set four times higher in the vertical direction than
that of horizontal direction. Model artifacts are sup-
pressed by a proper choice of the damping constant
(Van Avendonk et al. 2004). The overall misBt is
Figure 5. (a) The downward continuation (DC) result of the
measured by v2, which is the sum of the squared
shot gather located at location B (Bgures 1 and 2c) after individual travel time misBts divided by the overall
placing the source and receiver positions on the seaCoor. pick uncertainty. The misBt is gradually reduced
Downward continuation separates the superposition of refrac- through a series of linearized inversions until the
tion and reCection events and ensures that the refractions are desired Bt is achieved, typically v2 = 1, at which point
the Brst arrivals on the shot record (red) predicted and
the average misBt matches the pick uncertainty. At
(yellow) observed; (b) corresponding ray tracing. Ray fans
map until 3 km below seaCoor; and (c) starting misBt (blue) each iteration, velocity nodes are updated throughout
and Bnal misBt (red). the ray covered portion of the model.
44 Page 6 of 9 J. Earth Syst. Sci. (2021)130:44

(a) SW Distance (km) NE (b) SW Distance (km) NE


0 6 12
12 18 24 30 0 6 12
12 18 24 30

NMO
C Subduction Front
3

B
Depth (km)

4
A Trench

6
(c) (d)
OBS tomography
3
Subduction lope
tal s
Front Fron
Depth (km)

4
Trench

6
(e) (f)
DC + tomography
3
Depth (km)

6
V(km/s)
2.0 2.5 3.0 3.5 4.0 4.5

Figure 6. Velocity models on a 500-m grid and associated pre-stack depth migrations of the 5.5 km towed streamer data for
(a) interval velocities from the 5.5 km towed streamer data, (b) pre-stack depth migrated section using interval velocity derived
from NMO picking velocity, (c) tomographic inversion of sparse OBS data recorded at the sea-Coor, (d) pre-stack depth
migrated section using interval velocity derived from inverted OBS tomography velocity, and (e) tomographic inversion of
downward continued 12.0 km towed streamer data, and (f) pre-stack depth migrated section using interval velocity derived from
integrated technique. A, B and C indicate location of shot gathers shown in Bgure 2.

The ray tracing of this downward continued shot from NMO velocity analyses of 5.5 km streamer
gather (Bgure 5a), which is located on the frontal data, and the travel time tomography of OBS and
slope at location B (Bgure 1), is described in downward continued 12 km long streamer data.
Bgure 5(b). The shot and receivers are located on We have used the same short streamer (5.5 km)
the seaCoor after downward continuation and the data as used in the conventional processing, for
rayfans mapped the subsurface until a depth of *7 pre-stack depth migration wherein sampling
km (Bgure 5b). The travel time misBt is described interval was kept as 2 ms in order to obtain high-
in Bgure 5(c) in which the blue and red lines resolution image; in addition, preprocessing steps
demarcate the starting (*120 ms) and Bnal misBts were applied to remove swell noise attenuation
(*15 ms), respectively, between the observed and and multiples. Apart from this, we have also
the predicted travel times. applied inner trace mute on these data to eAec-
tively suppress the existing residual multiples
3.2.3 Pre-stack depth migration of 5.5 km from the near oAset. EAect of side scattering is
streamer data reduced by applying dip Bltering. Moreover, a
mild triangular anti-aliased Blter was used by
The interval velocity, which is required as the setting aperture length to 3.5 km during
input for pre-stack depth migration, was obtained migration.
J. Earth Syst. Sci. (2021)130:44 Page 7 of 9 44

(a) (b) (c)


A-5 km B-12 km C- 23 km

Depth (km)

Velocity (km/s) Velocity (km/s) Velocity (km/s)

Figure 7. 1D velocity proBles at A, B and C locations. (blue) NMO derived interval velocity, (yellow) OBS tomography derived
velocity, (red) downward continuation of streamer data derived velocity.

Distance (km) (1) interval velocities calculated from the 5.5 km


0 1.5 3.0 4.5 towed streamer data using conventional NMO
0
approach (Bgure 6a), (2) tomographic inversion of
the sparse OBS data using travel time tomographic
study (Bgure 6c) adapted from Singh et al. (2012),
and (3) the travel time tomographic inversion of
2 the downward continued 12.0 km towed streamer
data obtained from this study (Bgure 6e). The
interval velocity based on 12 km streamer data was
4 similar to that of 5.5 km streamer data owing to
Depth (km)

high slope. In each instance, the velocity–depth


model was interpolated to a 500-m grid and was
used for pre-stack depth migration of the 5.5 km
6
long towed streamer data and are shown in
Bgure 6(b, d and f).
A comparison among different 1D velocity–depth
8 proBles across the subduction front at A, B and
C locations for these three types of velocities are
described in Bgure 7. The NMO analysis derived
10
interval velocity shows a smoothed step like
function, whereas the OBS tomography derived
Figure 8. Common image gather (CIG) obtained from proBles shows a very smooth increase in velocity
pre-stack depth migration of 5.5 km streamer data using with depth. The high resolution downward
interval velocity of tomographic inversion of downward
continued velocity proBles demarcate the
continued 12 km long streamer data.
sharp change in the velocity gradients with
4. Results: Velocity model determination depth.
and pre-stack depth images The velocity–depth model from the tomographic
inversion of the OBS data produced poor migrated
Figure 6 (left panels) shows three velocity models image (Bgure 6d), simply because the OBSs were
that were determined from the data recorded over too far apart (*8.1 km) and their data could not
30 km segment of WG2 proBle along the slope: accurately characterise the velocities in the
44 Page 8 of 9 J. Earth Syst. Sci. (2021)130:44

topographically complex near surface. The interval the shallow subsurface structures and eventually
velocities calculated from the 5.5 km streamer data lead to poor depth migrated image.
are inaccurate and produce a poorly focused depth The best image of the accreted sediments below
migrated image (Bgure 6b). Tomographic inversion the frontal slope is obtained using the downward
of the 12.0-km long oAset streamer data produces continued streamer datasets. Relatively Bne shot
the most accurate velocity–depth model. The and group intervals (50 and 12.5 m, respectively)
resulting migrated image is sharp with clear defi- associated with a very long streamer (12 km)
nition of reCection events and faults (Bgure 6f). accurately imaged the sedimentary column until a
The common image gather (CIG) obtained after depth of 3 km below the frontal slope. As the
the pre-stack depth migration of the short streamer refracted arrivals are very sensitive to the Cuctu-
data using tomography derived velocity–depth ations of the Bne layers, the inversion results from
model of downward continued gathers is shown the downward continued data ultimately became
in Bgure 8 in which the reCection events are very useful for pre-stack depth migration providing
straightened justifying, further, the accuracy of the the best quality of subsurface image.
velocity–depth model. Although we have constrained our study to
pre-stack depth migration, the downward contin-
uation results can further be used for various other
5. Discussion and conclusions perspectives. For instance, the high-resolution
velocity model derived from the downward con-
Present study focuses on the eAect of interval tinued data can, further, be used for full waveform
velocity–depth model derived from three different inversion (Arnulf et al. 2012). By downward con-
modelling techniques: (1) conventional NMO based tinuing the surface seismic data, we can reduce the
velocity analysis, (2) travel time tomography of computation time for waveform inversion signifi-
sparsely (8.1 km apart) deployed OBSs, and (3) cantly. Since the rays for the Brst arrival refractions
travel time tomography of downward continued travel horizontally, they provide information on
streamer data. The study area is located at a horizontal velocity whereas reCection data provide
complex geological setting, i.e., the frontal slope of information on vertical velocity.
northern Sumatra subduction system, where the
water depth varies significantly from the trench to
the slope region and bathymetry is very rough. Acknowledgements
Moreover, the sedimentary deposits are also con-
solidated updip away from the trench, further We are thankful to J Martin, A Harding, N Bangs,
intricating the acquired seismic data. Due to the A Chauhan, H Carton and N D Hananto for their
presence of hard and crystalline accretionary support during this study. We are grateful to
wedge material, it is hard for seismic energy to Western Geco for carrying out the survey for us.
penetrate deeper. Since the conventional NMO
analysis is based on the approximation of Cat sur-
face or shallow dipping reCectors only, the sudden Author statement
changes of dip along the frontal slope cause severe
complicacies in carrying out velocity modelling Dibakar Ghosal contributed in writing, generating
using this method and ends up with an erroneous Bgures and editing. Satish Singh contributed in
velocity–depth model, which eventually leads to providing datasets and supervising the project.
inaccurate imaging of the subsurface folded and
faulted beddings. References
The OBS tomography could be a great alterna-
tive to this NMO based technique, but due to Arnulf A F, Singh S C, Harding A J, Kent G M and Crawford
limited number of OBSs with sparse spacing W 2011 Strong seismic heterogeneity in layer 2A near
between two consecutive OBSs, as the actual hydrothermal vents at the Mid-Atlantic Ridge; Geophys.
Res. Lett. 38 L13320.
objective of the survey was to model the nature of
Arnulf A F, Harding A J, Singh S C, Kent G M and Crawford
the whole subducting slab (Chauhan 2010; Singh W 2012 Fine-scale velocity structure of layer 2A from full
et al. 2012) across the trench, the spatial and waveform inversion of downward continued seismic reCec-
temporal resolution remained in the kilometer scale tion data in the Lucky strike volcano, Mid Atlantic Ridge;
and was not eDcient to model the Bner details of Geophys. Res. Lett. 39 8.
J. Earth Syst. Sci. (2021)130:44 Page 9 of 9 44
Berryhill J R 1979 Wave equation datuming; Geophysics 44 Liu C S, Curray J R and McDonald J M 1983 New constraints
1329–1344. on the tectonic evolution of the eastern Indian Ocean; Earth
Chauhan A P S, Singh S C, Hananto N D, Carton H, Planet. Sci. Lett. 65 331–342.
Klingelhoefer F, Dessa J, Permana H, White N J and McCaffrey R 2009 The tectonic framework of the Sumatra
Graindorge D 2009 Seismic imaging of forearc backthrusts subduction zone; Ann. Rev. Earth Planet. Sci. 37
at northern Sumatra subduction zone; Geophys. J. Int. 179 345–366, https://doi.org/10.1146/annurev.earth.031208.
1772–1780. 100212.
Chauhan A J 2010 Structure of the Northern Sumatra subduc- Matson R and Moore G F 1992 Structural controls on forearc
tion megathrust using seismic reCection and refraction data, basin subsidence in the central Sumatra forearc basin; Geol.
PhD Thesis, Institut de Physique du Globe de Paris. Geophys. Cont. Margins, AAPG Memoir 53 157–181.
Claerbout J 1976 Fundamentals of Geophysical Data Process- Moser T J 1991 Shortest path calculation of seismic rays;
ing, http://sepwww.stanford.edu/sep/prof/. Geophysics 56 59–67.
Galloway W E 2008 Depositional evolution of the Gulf of SheriA R E and Geldart L P 1982 Exploration seismology;
Mexico sedimentary basin; In: The Sedimentary Basins of volume 1, Cambridge University Press, 253p.
the United States and Canada, Sedimentary Basins of the Sieh K and Natawidjaja D 2000 Neotectonic of the Sumatran
World (ed.) Hsu K J, Elsevier, The Netherlands 5 505–549. fault Indonesia; J. Geophys. Res. 105 28,295–28,326,
Gazdag J 1978 Wave equation migration with the phase shift https://doi.org/10.1029/2000JB900120.
method; Geophysics 43 1342–1351. Singh S C, Carton H, Tapponnier P, Hananto N, Chauhan A P
Ghosal D, Singh S C, Chauhan A P S and Hananto N 2012 S, Hartoyo D, Bayly M, Moeljopranoto S, Bunting T,
New insights on the oAshore extension of the Great Christie P, Lubis H and Martin J 2008 Seismic evidence for
Sumatran fault, NW Sumatra, from marine geophysical broken oceanic crust in the 2004 Sumatra earthquake
studies; Geochem. Geophys. Geosyst. 13, https://doi.org/ epicentral region; Nat. Geosci. 1 771–781.
10.1029/2012GC004122. Singh S C, Chauhan A P S, Calvert A J, Hananto N D,
Ghosal D, Singh S C and Martin J 2014 Shallow subsurface Ghosal D, Rai A and Carton H 2012 Seismic evidence of
morphtectonics of the NW oAshore northern Sumatra bending and unbending of subducting oceanic crust and
subduction system using an integrated seismic imaging the presence of mantle megathrust in the 2004 Great
technique; J. Geophys. Res. 198 1818–1831, https://doi. Sumatra earthquake rupture zone; Earth Planet. Sci. Lett.
org/10.1093/gji/ggu182. 321–322 166–176, https://doi.org/10.1016/j.epsl.2012.01.
Fitch T J 1972 Plate convergence, transcurrent faults, and 012.
initial deformation adjacent to southeast Asia and the Van Avendonk H J A, Shillington D, Holbrook W S and
western PaciBc; J. Geophys. Res. 77 4432–4460, https:// Hornbach M 2004 Inferring crustal structure in the Aleu-
doi.org/10.1029/JB077i023p04432. tian island arc from a sparse wide-angle seismic data set;
Harding A J, Kent G, Blackman D K, Singh S C and Cannales Geochem. Geophys. Geosyst. 5(8) 1527–2087.
J-P 2007 A new method for MCS refraction data analysis Yelisetti S, Spense G D and Riedel M 2014 Role of gas hydrate
of the uppermost section at a Mid-Atlantic Ridge core in slope failure of frontal ridge of northern Cascadia margin;
complex; EOS, Trans. Am. Geophys. Union 88(52), Fall Geophys. J. Int. 199(1) 441–458, https://doi.org/10.1093/
Meet. Suppl., Abstract S12A-03. gji/ggu254.

Corresponding editor: ARKOPROVO BISWAS

You might also like