Vous êtes sur la page 1sur 17

Research paper

Deep crustal structure of the conjugate margins of the SW South China


Sea from wide-angle refraction seismic data
T. Pichot
a,
*
, M. Delescluse
a
, N. Chamot-Rooke
a
, M. Pubellier
a,1
, Y. Qiu
b
, F. Meresse
a, 2
,
G. Sun
b
, D. Savva
a
, K.P. Wong
a, 3
, L. Watremez
c
, J.-L. Auxitre
d
a
Laboratoire de Gologie, Ecole normale suprieure, CNRS/UMR 8538, 24 rue Lhomond, Paris 75005, France
b
Guangzhou Marine Geological Survey, Guangzhou 510760, China
c
Department of Oceanography, Dalhousie University, P.O. Box 15000, Halifax, NS, B3H 4R2, Canada
d
Total PN/BTF e Geosciences New business, 2 Place Jean Millier, La Dfense Paris, France
a r t i c l e i n f o
Article history:
Received 3 July 2013
Received in revised form
10 September 2013
Accepted 16 October 2013
Available online xxx
Keywords:
South China sea basin
Continental extension
Crustal structure
Wide-angle refraction seismic
Seismic tomography
Moho
a b s t r a c t
The South China Sea is the largest marginal basin of SE Asia, yet its mechanismof formation is still debated.
A1000-kmlong wide-angle refractionseismic prole was recentlyacquiredalong the conjugate margins of
the SWsub-basin of the South China Sea, over the longest extended continental crust. Ajoint reection and
refraction seismic travel time inversion is performed to derive a 2-Dvelocity model of the crustal structure
and upper mantle. Based on this new tomographic model, northern and southern margins are genetically
linkedsince theyshare commonstructural characteristics. Most of the continental crust deforms ina brittle
manner. Two scales of deformation are imaged and correlate well with seismic reection observations.
Small-scale normal faults (grabens, horsts and rotated faults blocks) are often associated with a tilt of the
velocity isocontours affecting the upper crust. The mid-crust shows high lateral velocity variation dening
low velocity bodies bounded by large-scale normal faults recognized in seismic reection proles. Major
sedimentary basins are located above low velocity bodies interpreted as hanging-wall blocks. Along the
northern margin, spacing between these velocity bodies decreases from90 to 45 kmas the total crust thins
toward the ContinenteOcean Transition. The ContinenteOcean Transitions are narrow and slightly
asymmetric e60 kmon the northern side and no more than 30 kmon the southern side eindicating little
space for signicant hyper-stretched crust. Although we have no direct indication for mantle exhumation,
shallow high velocities are observed at the ContinenteOcean Transition. The Moho interface remains
rather at over the extended domain, and remains undisturbed by the large-scale normal faults. The main
dcollement is thus within the ductile lower crust.
2013 Published by Elsevier Ltd.
1. Introduction
The South China Sea (SCS) is one of the largest submerged
continental provinces of SE Asia, covering a surface of nearly
1500 1000 km. Convergence of the Pacic plate toward Eurasia
during the Mesozoic and the Cenozoic times has shaped the region,
plate rollback leading to the progressive dislocation of the South
China continent (known as South China Block) and the formation of
a series of marginal basins at the edge of Sundaland. Whereas Sulu
Sea and Celebes Sea opened as back-arc basins, the origin of the SCS
involves the formation of a proto South China Sea probably oored
with oceanic crust that has now been subducted (Pubellier et al.,
2003). The V-shaped morphology of the SCS results from the
mid-Neogene propagation toward the SW of a now extinct NE-SW
spreading center (Taylor and Hayes, 1983; Briais et al., 1993;
Huchon et al., 1998, 2001).
Bathymetry and gravity data show complex longitudinal
morphologic variations along the entire SCS province. While the
eastern part of the SCS northern margin exhibits 400 km of
extended crust, its western part shows nearly 800 km of extended
continental crust (Fig. 1) which makes it one of the widest rifted
margin in the world. This variability can be partly attributed to the
heterogenous nature of the crust of the South China Block, in
relation to a complex evolution of accreted terranes of Gondwana
* Corresponding author. Present address: Institut de Physique du Globe de Paris,
Sorbonne Paris Cit and CNRS/UMR 7154, 1 rue Jussieu, F-75238 Paris CEDEX 05,
France.
E-mail address: thibaud.pichot@gmail.com (T. Pichot).
1
Present address: Faculty of Geosciences and Petroleum Engineering, Universiti
Teknologi Petronas, 31750 Tronoh, Perak Darul Ridzuan, Malaysia.
2
Present address: Total SA, Structural Geology Group, CSTJF, avenue Larribau,
64000 Pau, France.
3
Present address: Department of Geosciences, University of Oslo, Norway.
Contents lists available at ScienceDirect
Marine and Petroleum Geology
j ournal homepage: www. el sevi er. com/ l ocat e/ marpet geo
0264-8172/$ e see front matter 2013 Published by Elsevier Ltd.
http://dx.doi.org/10.1016/j.marpetgeo.2013.10.008
Marine and Petroleum Geology xxx (2013) 1e17
Please cite this article in press as: Pichot, T., et al., Deep crustal structure of the conjugate margins of the SW South China Sea from wide-angle
refraction seismic data, Marine and Petroleum Geology (2013), http://dx.doi.org/10.1016/j.marpetgeo.2013.10.008
origin since the Phanerozoic time (Pubellier et al., 2003),
including Kwangsian (middle Paleozoic), Indosinian (Triassic),
and Yanshanian (JurassiceCretaceous) phases (Wang et al., 2012).
The age and distribution of JurassiceCretaceous granites along the
South China Block and the present-day southern margin of the
SCS (including Dangerous Grounds and Palawan blocks) attrib-
uted to the Yanshanian tectonic phase advocate for an Andean-
type arc inheritance. This arc was associated with the
northwestward-dipping subduction of the Paleo-Pacic plate
below the South China Block during the Middle JurassiceMiddle
Cretaceous (Jahn et al., 1976; Schlter et al, 1996; Yan et al., 2010;
Wang et al., 2012).
Longitudinal contrast along the northern margin suggests that
different modes of extension prevailed along the continental mar-
gins of the SCS (Hayes and Nissen, 2005). Today, fully dynamic
models using non-linear thermo-mechanical properties are able to
reproduce the great diversity of styles of observed structures at
continental margins by introducing depth-dependent rheology and
complex combination of pure (McKenzie, 1978) and simple shear
(Wernicke, 1985) deformation in the crust and the mantle
(Huismans and Beaumont, 2003; Huismans et al., 2005; Lavier and
Manatschal, 2006; Huismans and Beaumont, 2011). This led to
more sophisticated models such as polyphase faulting, sequential
faulting, crustal embrittlement and depth-dependent stretching
(Reston, 2005; Ranero and Prez-Gussiny, 2010; Prez-Gussiny
et al., 2003; Davis and Kusznir, 2004; Lavier and Manatschal,
2006; Huismans and Beaumont, 2011).
In the light of these recent advances, the SCS is an exceptional
natural laboratory. It makes an interesting area to study the cause of
lateral discrepancies in structural style found along SCS rifted
margins during rifting. Although the northern margin of the SCS
was intensively studied, the western segment of the SCS that rep-
resents the widest area of extended continental crust remains
poorly explored. A recent 1000-km-long wide-angle refraction
seismic prole has been acquired in this least studied part of the
SCS province, the conjugate margins of the SWSCS (Fig. 1). Here, we
perform a joint reection and refraction seismic travel times
inversion to build a 2-D velocity model of the thinned continental
crust of the SW conjugate margins. We analyze the links between
the distribution of P-wave velocities with depth and the petrolog-
ical layering of the continental crust. We then suggest possible
implications in terms of rheological behavior of the continental
crust under the extension processes involved during the opening of
the SW SCS.
2. Data acquisition, processing and modeling
2.1. Data acquisition
In June 2011, Chinese and French scientists from the Guangzhou
Marine Geological Survey and from the Geological Laboratory of
cole Normale Suprieure conducted a joint geophysical experi-
ment on board the R/V Tan Bao across the SW sub-basin of the SCS.
The experiment consisted in the acquisition of a w1000-km-long
Figure 1. The South China Sea (SCS). A) Bathymetric map (Smith and Sandwell, 1997) showing the main geological provinces and countries adjacent to the SCS. The thin black line
indicates the Continent-Ocean Transition (COT). The heavy dashed line corresponds to the SCS spreading center (from Briais et al., 1993). The colored stars represent the locations of
dredged samples. Associated lithologies and ages (from Kudrass et al., 1986; Yan et al., 2010) are displayed at the bottom of the gure. B) Bathymetric map contoured every 1000 m
showing the main refraction proles acquired in the SCS, ESP-W1985; ESP-C1985; ESP-E1985 (Nissen et al., 1995); OBS 1993 (Pin et al., 2001); OBH 1996-IV (Qiu et al., 2001); OBS
2001 (Wang et al., 2006); OBS 2006-1 (Wu et al., 2012); OBS 973-1 (Qiu et al., 2011); OBS973-2 (Ruan et al., 2011); OBS 973-3 (Lu et al., 2011). Interpreted magnetic anomalies shown
in thin black lines are from Briais et al. (1993). A red line indicates the shot positions of the OBS prole presented in this study, while small black dots display the OBS positions. (For
interpretation of the references to color in this gure legend, the reader is referred to the web version of this article).
T. Pichot et al. / Marine and Petroleum Geology xxx (2013) 1e17 2
Please cite this article in press as: Pichot, T., et al., Deep crustal structure of the conjugate margins of the SW South China Sea from wide-angle
refraction seismic data, Marine and Petroleum Geology (2013), http://dx.doi.org/10.1016/j.marpetgeo.2013.10.008
Ocean Bottom Seismometers (OBS) prole across the conjugate
margins and the oceanic basin. The section runs NW-SE perpen-
dicular to the inactive spreading center axis and a total of 50 OBSs
were deployed along the line (Fig. 1). Twenty-four Sercel Microbs

OBSs were deployed over about half the prole, covering the
Spratly Islands area (Dangerous Grounds province), the oceanic
domain and the distal portion of the northern conjugate margin
(i.e. OBS-1 to OBS-24). The remaining OBSs were provided by the
Institute of Geology and Geophysics, Chinese Academy of Sciences,
and cover the rest of the northern margin (i.e. OBS-25 to OBS-50).
For technical reasons, the originally planned seismic reection
line across the set of OBSs could not be shot, so that previous
seismic lines around were further used to interpret the refraction
results, in particular the sub-surface structures.
The rationale of the experiment was to focus as much as possible
on the conjugate continental domains and transition zones, using a
reasonable number of OBSs. The mean distance between OBSs was
thus kept close to 18 km in the continental parts (Fig. 2). A larger
and variable spacing was used in the oceanic domain, which was
not the primary goal of the study. Thus, oceanic crustal velocities
were obtained with fewer details.
All OBSs were recovered successfully and only two did not re-
cord any data (OBS-28 and OBS-50). The source was designed to
investigate deep crustal structures and consisted of an array of Bolt
air guns with a total volume of w105 l. Shots were distance trig-
gered every 150 m (leading to a 1 min trace length). A total of ca.
7000 shots were recorded on each OBS. The sample interval used
was 1, 4 or 8 ms, depending on the OBS type. All records were
resampled to 4 ms for the nal processing. The OBS records are of a
good overall quality (Fig. 3). For each instrument, the hydrophones
and the vertical component of the geophones were used for seismic
phase identications. Clock drifts were checked and corrections
were applied. A relocation procedure was applied using water wave
travel times and multibeam bathymetry. On average, relocated
positions of the OBSs showed a drift of the order of w300 m along
inline and/or cross line directions.
2.2. Seismic phase identication
We manually picked a total of 41,100 rst refraction arrivals (Pg
phases, rays refracted in the sediments or the crust and Pn phases,
rays refractedinthe upper mantle). Pnphases weredetectedas far as
200 km offset in many cases (Fig. 3A) and occasionally beyond. A
picking error ranging from30 ms to125 ms was assigned depending
on the signal to noise ratio, using the empirical parameterization of
Zelt andForsyth(1994). A200ms pickingerror was attributedtoless
certain Pg and Pn picks. Most OBSs did not show any sedimentary
second arrival, while at near offset, the rst refraction arrivals most
often showed too high of an apparent velocity to imply a thick
sedimentary pile (Fig. 3A). However, where sedimentary basins do
exist (mainly on the northern margin) refraction phases from the
sediments are rst arrivals (Fig. 3B and C).
PmP phases (i.e. reection phases from the crust-mantle
boundary) were identied on most of the OBSs deployed above
the continental crust (Fig. 3) and correspond to a set of 6622 picks.
A 150 ms picking error was assigned for this reection phase. No
PmP arrival was recognized in the oceanic basin (Fig. 3D). For OBSs
located close to the continental slope, Pg arrivals display a strong
asymmetrical pattern suggesting a sharp transition between the
oceanic crust and the continental crust in termof velocity structure
(Fig. 3E). In most cases where PmP arrivals were picked, the Pg-Pn
crossover was recognized (known as triplication point) providing
better constrains for the determination of the Moho depth (Fig. 3C).
Some OBS data display a Pg phase at very far offset where the Pn
phase is already clearly visible, suggesting a strong lateral velocity
variation within the continental crust (Fig. 3F).
Figure 2. A) Multibeam bathymetric map along the OBS prole (see Fig. 1 for location). B) Free air gravity anomaly map (Sandwell and Smith, 2009). The positions of the OBS are
reported in yellow dots. The red open circles correspond to OBS that did not record any data. The magnetic anomalies interpreted by Briais et al. (1993) are indicated in thick dashed
lines. Bathymetric (C) and gravimetric (D) cross-sections along the OBS prole. Note that the regional trend of the free air gravity anomaly along most of the OBS prole is zero. (For
interpretation of the references to color in this gure legend, the reader is referred to the web version of this article).
T. Pichot et al. / Marine and Petroleum Geology xxx (2013) 1e17 3
Please cite this article in press as: Pichot, T., et al., Deep crustal structure of the conjugate margins of the SW South China Sea from wide-angle
refraction seismic data, Marine and Petroleum Geology (2013), http://dx.doi.org/10.1016/j.marpetgeo.2013.10.008
2.3. Modeling procedure
Different 2-D velocity modeling strategies are used for wide-
angle seismic data: forward modeling (Zelt and Smith, 1992), in-
verse modeling (Zelt and Barton, 1998; Van Avendonk et al., 1998;
Korenaga et al., 2000) or a combination of both (Zelt and Smith,
1992). We use here the joint inversion method of Korenaga et al.
(2000) that combines refraction and reection travel times. The
main reason for choosing Korenagas inversion scheme is that it
does not require strong a priori information, apart from a loose
initial velocity model. This is particularly important in our case
since we did not collect the seismic reection line. Inverting for the
depth of a oating reector (here, the Moho interface) is also
possible. The rationale is to provide a simple model that ts the
data with little risk of pre-inversion over-interpretation. All
methods (forward or inverse modeling) produce non-unique re-
sults. As such, we will later discuss the preferred results in the light
of what is known of the geology of the area.
The velocity model is parameterized as a two-dimensional
sheared mesh grid. To allow for a higher level of sub-surface de-
tails, vertical grid cells size increases from 200 m at the top to
500 m at the bottom of the model, with a constant horizontal
length of 500 m. The multibeambathymetric data denes the top of
the sheared grid. Because no seismic reection data has been ac-
quired collectively with the refraction seismic experiment, the
interface between sediments and the basement has not been
introduced in the initial velocity model. As stated in Section 2.2,
sedimentary refracted arrivals are always rst arrivals at the
Figure 3. Receiver-gather of OBS data using the vertical component of the geophone (top). First refracted (red) and reected (green) arrival picks with their corresponding error
bars (middle). Observed (colors) and modeled travel times (black dots) using the nal velocity model presented in Figure 4 (bottom). The vertical axis is in reduced travel time and
the horizontal axis represents the offset from the OBS position (in kilometers). The reduced travel time is expressed as T-offset/(reduced velocity). A refraction phase with an
apparent velocity higher than the chosen reduced velocity would have travel times decreasing with offset, while a phase with a lower apparent velocity would have travel times
increasing with offsets. (For interpretation of the references to color in this gure legend, the reader is referred to the web version of this article).
T. Pichot et al. / Marine and Petroleum Geology xxx (2013) 1e17 4
Please cite this article in press as: Pichot, T., et al., Deep crustal structure of the conjugate margins of the SW South China Sea from wide-angle
refraction seismic data, Marine and Petroleum Geology (2013), http://dx.doi.org/10.1016/j.marpetgeo.2013.10.008
continental domain water depths. As a result, the inversion should
retrieve the large sedimentary basins without difculty.
A Bouguer gravity inversion has been performed to dene a
crude Moho interface to be used as input in the initial velocity
model. The inversion procedure is using Parkers Fourier formula-
tion for gravity (Parker, 1973) widely applied in sedimentary basins
(details of our procedure can be found in Chamot-Rooke et al., 1997;
see Braitenberg et al., 2006; Nguyen and Nguyen, 2013 for the latest
applications to the SCS). Since the aim was to obtain a smooth
starting velocity model, we did not try to introduce sophisticated
corrections such as variable crustal densities due to the nature of
the crust (continental versus oceanic) or variable mantle density
due to the thermal anomaly related to the SCS opening. A standard
2700 kg/m
3
crustal density was used to rst derive a Bouguer
anomaly based on the multibeam bathymetry. The resulting
anomaly was then converted to a Moho geometry using a crust/
mantle contrast of 450 kg/m
3
. The reference depth for the inversion
was set to 20 km, so that the gravity inverted oceanic crust thick-
ness was kept close to 6 km, and the depth of the continental Moho
interface was in good agreement with previous seismic refraction
studies (Nissen et al., 1995; Pin et al., 2001; Qiu et al., 2011; Ruan
et al., 2011; L et al., 2011). The initial velocity model is
composed of a linear velocity gradient from 4 km/s below the
seaoor interface to 7.5 km/s at the crust-mantle boundary (a ve-
locity of 7.5 km/s is quite high but this avoids a strong velocity
contrast at the crust-mantle boundary in the initial velocity model).
Avelocity of 8 km/s is applied in the shallowupper mantle (Fig. 4A).
The inversion method used here requires a large set of param-
eters (detailed in Table 1; see also Korenaga et al., 2000). The key-
parameters are summarized hereafter. Both velocity and depth
perturbations are controlled by predened vertical and horizontal
correlations lengths (Toomey et al., 1994). After exploring a large set
of parameters, the preferred values for the correlation lengths are
10e30 km horizontally and 0e24 km vertically, giving a value of
Figure 3. (continued).
T. Pichot et al. / Marine and Petroleum Geology xxx (2013) 1e17 5
Please cite this article in press as: Pichot, T., et al., Deep crustal structure of the conjugate margins of the SW South China Sea from wide-angle
refraction seismic data, Marine and Petroleum Geology (2013), http://dx.doi.org/10.1016/j.marpetgeo.2013.10.008
several hundreds meters at the seaoor and nearly 12 km at 20 km
depth (the rst value is applied to the top of the grid, the second is
applied to the bottom, while linear interpolation is performed in
between).
To minimize the computation time and the memory cost of the
runs, the w1000-km-long prole was divided into two parts, a 740-
km-long northern prole (PR1), and a 458-km-long southern pro-
le (PR2). The proles share a common oceanic domain around
OBS-18 (Fig. 5).
3. Results of the velocity modeling
3.1. Robustness of the results
As mentioned above, the result of the inversion is non-unique.
The nal velocity model presented in Figure 4B has the lowest
nal root mean square (RMS) residual and reduced c
2
among the
tested models and satises the rst order geology (discussed
later). After 12 iterations, the nal RMS residual is nearly equal for
both the northern (PR1) and southern (PR2) proles with a value
of w160 ms. Note that the average picking uncertainties is slightly
higher for PR1 than for PR2 because of a lower signal-to-noise
ratio for OBSs along PR1. The nal c
2
values are 1.55 and 2.02 for
PR1 and PR2, respectively (see Table 2 for details). The expected
reduced c
2
values should be close to 1 if the post-inversion travel
time residuals were close to the initial picking uncertainties.
However, this does not take into account the aliasing caused by
the w18 km OBS spacing. Considering that the basement interface
was not constrained by seismic reection data we explain values
of c
2
above 1 by unresolved basement topography. The higher c
2
for PR2 illustrates this, as it is likely due to the combination of
lower average uncertainties and higher amplitudes of seaoor and
basement topography at short wavelengths (i.e. Dangerous
Grounds).
Figure 3. (continued).
T. Pichot et al. / Marine and Petroleum Geology xxx (2013) 1e17 6
Please cite this article in press as: Pichot, T., et al., Deep crustal structure of the conjugate margins of the SW South China Sea from wide-angle
refraction seismic data, Marine and Petroleum Geology (2013), http://dx.doi.org/10.1016/j.marpetgeo.2013.10.008
For both proles PR1 and PR2, the refracted ray coverage (Pg
and Pn) is particularly dense in the rst 10 km (Fig. 5AeB). The
refracted ray density is slightly higher in PR2 than in PR1 below
10 km. Many refracted arrivals turn into the shallow upper mantle
(Pg) giving good constraints on the velocity at the base of the
model (Fig. 5AeB).
Reected rays (PmP) are traced along sizeable portions of the
proles: w500 km for PR1 (between km 30 and 540) and 200 km
for PR2 (between km 200 and 400; Fig. 5CeD). In particular, con-
tinental slopes adjacent to the oceanic domain show good wide-
angle reection ray coverage. Rays from several OBSs cover the
crust of the ContinenteOcean Transitions (COTs) on both conjugate
margins, showing a reasonably good coverage (Fig. 5). However, the
resolution of the velocity model is limited in these areas since
almost no crustal arrival is recorded as rst arrival (crustal arrivals
are rst arrivals only at a very short offset). As a result, the obtained
nal velocity gradient is similar to the initial velocity gradient.
Velocities are also poorly constrained in the central part of the
oceanic domain, refracted rays being recorded by OBS-18 only. As a
result, the obtained nal velocity there does not really improve
compared to the initial velocity gradient. Furthermore, no reection
phase was identied at the oceanic crust/mantle boundary.
Small-amplitude and small-wavelength undulations of the
Moho were found in many of the test runs. They are actually
superimposed onto more stable long-wavelength Moho variations.
These small-wavelength undulations may be artifacts related to the
Figure 3. (continued).
T. Pichot et al. / Marine and Petroleum Geology xxx (2013) 1e17 7
Please cite this article in press as: Pichot, T., et al., Deep crustal structure of the conjugate margins of the SW South China Sea from wide-angle
refraction seismic data, Marine and Petroleum Geology (2013), http://dx.doi.org/10.1016/j.marpetgeo.2013.10.008
difculty in the inversion to weight reected phases with respect to
refracted ones. If real, a plausible interpretation is that the Moho
discontinuity is not a sharp boundary, but rather a complex tran-
sition between the lower crust and the shallow upper mantle
(Korenaga and Kelemen, 1997; Ziegler, 2001). Seismic rays would
then possibly not be reected by a simple reector, but interact
with a heterogenous velocity gradient in the vicinity of the Moho, at
the base of the lower continental crust.
3.2. Velocity structures
Velocities directly below the seaoor are well below 4 km/s, the
assigned value there in the initial velocity model. They increase
rapidly with depth in the rst few kilometers from 1.6 km/s to 5.0e
6.0 km/s (Fig. 6AeB). Despite the absence of basement interface in
the initial velocity model, known major sedimentary basins and
basement highs are fairly well imaged. These basins are, fromNorth
to South, the Quiongdongnan Basin, the Guangle Uplift (also known
as the Triton Ridge), the Zhongjannan Basin (also known as the
Triton Lower Terrace), the southwestern area of the Penxi Sea-
mounts province and isolated basins and basement highs of the
Spratly islands (Fig. 4B). Narrow sedimentary basins are also
modeled at the feet of the continental slopes of both margins. The
largest of these basins is at the foot of the northern margin.
Within the upper 5 km, 1-D velocity-depth proles (Figs. 6DeE)
show larger variations over PR1 when compared to PR2. This is
essentially due to the presence of two large sedimentary basins on
the northern margin, the Quiondongnan Basin (tr1; Figs. 4B and
6D) and the Zhongjannan Basin (tr3; Figs. 4B and 6D). In the upper
section (rst 5 km), the 1.6 km/s to 5 km/s velocity isocontours
showsmall perturbations with an associated wavelength bracketed
between 15 and 30 km(Fig. 4B). This wavelength is probably not an
artifact since it is signicantly greater than the scaling parameters
used in the inversion, the vertical and horizontal correlation
lengths at shallow depth being much smaller. The wavelength of
these shallow perturbations varies along the prole. This is not
Figure 3. (continued).
T. Pichot et al. / Marine and Petroleum Geology xxx (2013) 1e17 8
Please cite this article in press as: Pichot, T., et al., Deep crustal structure of the conjugate margins of the SW South China Sea from wide-angle
refraction seismic data, Marine and Petroleum Geology (2013), http://dx.doi.org/10.1016/j.marpetgeo.2013.10.008
simply due to the velocity perturbation inherited from the seaoor
topography propagating further down into the velocity model.
Indeed, in the Zhongjannan Basin, between OBS-41 and OBS-36, the
upper part of the sedimentary cover (1.6e4.0 km/s) does not pre-
sent such perturbation while, velocity isocontours just below
(ranging between 4.0 and 5.5 km/s) show clearly the small wave-
length perturbation (Fig. 4B). The same wavelength can be seen in
the topography of the seaoor and in the free-air gravity anomaly
(especially along PR2; Fig. 2). In the discussion, we suggest that
these structures relate to the uppermost crust horst-and-graben
small-scale fabric.
A rather sharp change in the velocity gradient occurs in all
proles at w5.0e6.0 km/s (Fig. 5AeB). At intermediate depth (be-
tween 5 and 15 km below the sea level), the nal velocity model
shows strong lateral variation for velocities ranging between
w5.7 km/s to less than 6.7 km/s. Within this velocity range, the
entire prole, including PR1 and PR2, shows alternating bodies of
high and low velocity (Fig. 4B). The Low Velocity Bodies (LVBs)
show smaller velocity values when compared to the velocity
average of both PR1 and PR2 (compare Fig. 6FeG to Fig. 6H) and to
the crustal velocity model of the average continental margin type
(Fig. 6FeG) published by Mooney et al. (1998). Six LVBs were
identied along the northern margin and two more along the
southern margin (Fig. 4B). These bodies are not regularly spaced:
along PR1, the spacing reaches w90 km to the North and decreases
to a value of 45 km toward the oceanic crust. Along PR2, spacing is
about 70 km. Velocities higher than 6.5e6.7 km/s show the same
type of undulations although with far less amplitude.
Thin (2e3 km) high velocities lenses (7.0e7.7 km/s) are imaged
in some places within the lowermost crust above the Moho inter-
face (Figs. 4B and 6FeG). Because they are thin, we call them HVL
for High Velocity Lenses rather than HVB (High Velocity Bodies).
Along the northern margin (PR1), the largest of these lenses reaches
180 km between 130 and 310 km. A HVL is also present between
390 and 500 kmbut to a less extent (w110 km). Along the Southern
margins (PR2), a HVL is recognized between LVBs A and B over a
Figure 3. (continued).
T. Pichot et al. / Marine and Petroleum Geology xxx (2013) 1e17 9
Please cite this article in press as: Pichot, T., et al., Deep crustal structure of the conjugate margins of the SW South China Sea from wide-angle
refraction seismic data, Marine and Petroleum Geology (2013), http://dx.doi.org/10.1016/j.marpetgeo.2013.10.008
length of about 80 km (Figs. 4B). At both margins, HVL are also
present at the feet of the continental slope. Although relatively
limited on the northern side (w50 km in length, between 540 and
590 km), the southern HVL seems to be continuously present from
the necking zone to the COT (between 840 and 740 km; Fig. 4B).
The Moho interface is relatively at over the entire continental
margins (according to the vertical exaggeration; Fig. 4B). Along PR1,
the Moho gently rises southeastward with an average depth below
the sea level of w21 km (between 30 and 300 km) to w19 km (at
500 km). This change in depth of the Moho interface is marked by a
more pronounced rise around w400 km. This relative shallowing
coincides with the northeastern extent of the Phu Khan Basin (see
Figs. 1 and 4B). Along PR2, the Moho depth interface increases
slightly to the northwest from 15 km to 19.5 km bellow sea level.
The mean crustal thickness varies between 17.5 km and 9 km along
the northern margin (excluding the Quiongdongnan Basin) and
from 15 km to 8.5 km, along the southern margin (Fig. 7). The
thickest sediment thickness (16e17 km) is found within the
Quiongdongnan Basin, the crust below being reduced to w5 km.
Sharper changes occur at the COTs, the Moho rising from w19 km
to 10 km for both PR1 and PR2. The crust abruptly thins over a
distance of w100 km along PR1 (between 490 and 590 km) giving
an angle of 5

for the Moho interface. This distance does not exceed


60 km along PR2 (between 750 and 810 km; Fig. 4B) giving a
steeper slope of 8.5

.
In summary, the nal velocity model shows strong lateral ve-
locity variations within the intermediate depth continental crust
where low velocity bodies are imaged with a spacing ranging from
90 to 45 km. In a few places, high velocity lenses are imaged within
the lowermost crust, including at the ContinenteOcean Transition.
The Moho interface remains rather at over the entire continental
margins (except toward the ContinenteOcean Transition). Despite
a number of small differences between the two margins (mean
Figure 4. Preferred model for the entire OBS prole (i.e. PR1 and PR2 are merged). The positions of each OBS are reported in red. A) Initial velocity model with a linear velocity
gradient ranging from 4 km/s at the seaoor (thin black line) to 7.5 km/s at the Moho interface (heavy white line). The seaoor is extracted from multibeam bathymetry data and the
Moho interface is derived from Bouguer anomaly inversion. B) Final velocity model after 12 iterations. The heavy white line corresponds to the Moho interface constrained by PmP
arrivals. The shadow zone represents areas without ray coverage. The velocity isocontours are drawn every 0.25 km/s. The black arrows indicate the position of the 1-D velocity-
depth models presented in gure 6. The main geological provinces are annotated. Low Velocity Bodies (LVBs) are indicated by small number along the northern margin and by
letters along the southern margin. (For interpretation of the references to color in this gure legend, the reader is referred to the web version of this article).
Table 1
Tomo2D set of parameters used in this study.
Parameters Symbols Values used in
this study
Vertical correlation length for
velocity nodes
L
Vv
a
CV
b
10e30
Horizontal correlation length for
velocity nodes
L
Hv
a
CV
b
0e24
Correlation length for reector L
d
a
CD
b
7
Velocity smoothing l
v
a
SV
b
10
Depth smoothing l
d
a
SD
b
10
Velocity damping D
v
a
TV
b
10
Depth damping D
d
a
TD
b
e
Depth kernel weighting w
a
W
b
0.2
a
Refers to Korenaga et al. (2000).
b
Refers to Korenaga (2003).
T. Pichot et al. / Marine and Petroleum Geology xxx (2013) 1e17 10
Please cite this article in press as: Pichot, T., et al., Deep crustal structure of the conjugate margins of the SW South China Sea from wide-angle
refraction seismic data, Marine and Petroleum Geology (2013), http://dx.doi.org/10.1016/j.marpetgeo.2013.10.008
crustal thickness, Moho slope at the COT, spacing between the low
velocity bodies, size of the sedimentary basins), the average 1-D
velocity-depth models for each continental margin (PR1 and PR2)
are similar (Fig. 6C) suggesting a common origin.
4. Discussion
In the following section, the term northern margin and southern
margin will be preferred rather than PR1 and PR2, respectively.
4.1. Sub-surface crustal structure (sediments and upper to middle
crust)
The 1.6e5.5 km/s velocity range is consistent with seismic ve-
locities of the CenozoicePaleozoic sedimentary sequence
(including compacted sediments) reported in previous refraction
studies constrained by seismic reection proles in the SCS area
(Table 3).
The crustal structure that we nd in the uppermost part of our
section can be compared with what has been documented in the
seismic reection proles nearby (Hayes et al., 1995; Schlter et al.,
1996; L et al., 2011; Ding et al., 2013). Seismic reection data
collected in the vicinity of both conjugate margins actually show
subsurface extensional features commonly found at other rifted
continental margins such as small-scale grabens and horsts. In the
SCS, the dominant spacing between these small-scale subsurface
features is between 15 and 30 km (Hayes et al., 1995; Ding et al.,
2013). These structures, distributed over a wide area, are well
imaged at the seaoor and in the free-air gravity, so that the area
has been commonly referred to as a Basin-and-Range type of
rifted basin (Gilder et al., 1991; Zhang et al., 2009; Franke, 2012).
These small-scale normal faults are often associated with a tilt of
the velocity isocontours, not only through the entire sedimentary
cover, but also within the upper crust (Figs. 4B and 8).
Lateral velocity variations are highest at intermediate depth (5e
15 km), in the 5.7 to 6.5e6.7 km/s velocity range (Fig. 4B, variations
highlighted by the 6 km/s velocity contour). Nearby seismic
reection proles show the presence of localized large-scale deep-
rooted faults cutting through the entire crust (called through-going
crustal faults in Hayes et al., 1995) and controlling major tilted
faults blocks (Hayes et al., 1995; Hutchison and Vijayan, 2010; Ding
et al., 2013). These structures are consistent with the identied Low
Velocity Bodies (LVBs) as well as the spacing between them on the
nal velocity model (45e90 km). Along the northern margin, these
large-scale normal faults dip with an angle between 25

and 40

and each of them may have accommodated 5e20 km of horizontal


extension (Hayes et al., 1995). From the nal velocity model, it
appears that these faults dip equally north and south at both con-
jugate margins, as imaged by seismic reection proles acquired at
the southern margin (Schlter et al, 1996; Ding et al., 2013).
Both reection seismic data (Hayes et al., 1995; Ding et al., 2013)
and our nal velocity model suggest an average vertical displace-
ment on these large-scale normal faults of a few kilometers (Fig. 8).
The thick sedimentary basins are always located above the LVBs,
which are bounded by the hanging-walls of the large-scale normal
faults (Fig. 8). LVB A (Fig. 8) differs in its rather complex structure
and may correspond to a granitic massif according to the igneous
rocks sampled at the top of this structure (Yan et al., 2010) and its
relative homogenous velocity structure (Figs. 4B and 8).
4.2. A hint at the lower crustal rheology of the SCS conjugate
margins
Upper to mid-crustal velocity can be easily interpreted along
with available reection seismic proles. In most cases, the lower
crust is however more difcult to image using standard reection
imaging. This is where we have to rely solely on our refraction
velocity model. P-wave velocities from wide-angle refraction
measurement in the continental crust can be correlated to the
Figure 5. Compilation of the nal velocity model of PR1 (left) and PR2 (right) showing Pg and Pn (A and B), PmP (C and D) ray paths. Only one out of 2 rays is shown. Refer to
Figure 4 for more details. (For interpretation of the references to color in this gure legend, the reader is referred to the web version of this article).
T. Pichot et al. / Marine and Petroleum Geology xxx (2013) 1e17 11
Please cite this article in press as: Pichot, T., et al., Deep crustal structure of the conjugate margins of the SW South China Sea from wide-angle
refraction seismic data, Marine and Petroleum Geology (2013), http://dx.doi.org/10.1016/j.marpetgeo.2013.10.008
velocities of different rock types measured in laboratory
(Christensen and Mooney, 1995; Rudnick and Fountain, 1995;
Mooney et al., 1998). The difculty is that different petrology may
at the end lead to quite similar seismic velocities. Reston (2009)
pointed out that in the case of a hyper-extended lower crust
brought to the surface by polyphase faulting, the nal velocity of
the highly fractured rocks might actually resemble that of upper
crust material. However, we tentatively interpreted the velocities
obtained from our nal velocity model in terms of generic crustal
layers (here, upper crust and lower crust). Crustal velocities ranging
between 5.7 and 6.4 km/s are commonly measured in igneous and
metamorphic rocks of the upper and middle crust, such as diorites,
granites and paragneiss (Christensen and Mooney, 1995), whereas
velocities higher than 6.4 km/s are generally considered as lower
crustal material.
Abrupt lateral velocity variations at intermediate depth suggest
a brittle behavior of the mid-crust allowing the juxtaposition of
blocks with contrasting velocities (i.e. LVBs). Continental rocks
dredged in the Dangerous Grounds area were collected along the
footwalls of some of the largest half-grabens (Kudrass et al., 1986;
Yan et al., 2010; see Fig. 8). The dredged rocks mainly consist of
metamorphic gneiss and schists of Cretaceous age (although older
Jurassic metamorphic and Triassic shales have also been dredged).
They were exposed at the seaoor following the latest stretching
phase that opened the SCS, but their exhumation path remains
unknown since they may have been close to the surface earlier in
their history. In any case, the result is a contrasting velocity across
the large-scale normal faults, the contrast dying with the pro-
gressive attening of the large-scale normal faults as it roots in
horizontally sheared or owing material. Our interpretation is that
this rooting occurs deep into the lower crust. Velocities higher than
6.5e6.7 km/s show a smooth (i.e. almost at) geometry suggesting
ductile shear or ow. Note that not the entire lower crust seems to
deform in a ductile regime but just its lowermost part. A quartz-
dominated rheology under a thermal regime adequate for the
South China Sea predicts a brittleeductile transition at a depth of
10e15 km (Zuber et al., 1986; Clift et al., 2002). The ductile lower
crust is thus a natural candidate to act as a dcollement layer where
large-scale normal faults root. This result differs from previous
studies in the sense that the large-scale normal faults do not seem
to cut through the entire continental crust as suggested by Hayes
et al. (1995) or to root at the crust-mantle boundary as proposed
by Ding et al. (2013).
Flow within the lowermost crust is suggested by several prop-
erties of the nal velocity model. The lowermost crust is thin or
even absent beneath the thickest sedimentary basins, such as
belowLVBs 1, 3, and B (Figs. 4B and 7). This is best explained by ow
in the lowermost crust with a sense of shear opposite to the motion
on the large-scale normal fault. This results in a thicker accumu-
lation of lower crust at the base of the footwalls of the large-scale
normal faults as seen below LVB 1 (at 70 km), northeast of LVB 4
(at 400 km), and on both sides of LVB B (at 900 and 970 km; Figs. 4B
and 7).
Finally, large-scale normal faults bounding LVBs create mid-
crustal lateral velocity variations of signicantly higher ampli-
tudes than the depth variations of the Moho just below these
structures. The Moho remains rather undisturbed, as expected if
the dcollement is located within the lower crust. One consequence
is that at the size of the main LVBs, the isostatic compensation is
achieved in the lower crust. The corollary is that standard Moho
gravity inversions are of little interest for the understanding of the
mechanisms of extension.
4.3. Geophysical characterization of the continent-ocean transition
(COT)
The exact nature of the COT and its geometry remain poorly
constrained. The position of the COT in the SCS has been indirectly
identied using magnetic anomaly (Taylor and Hayes, 1983; Briais
et al., 1993; Barckhausen and Roeser, 2004; Li and Song, 2012) and/
or gravity anomaly (Taylor and Hayes, 1983; Braitenberg et al.,
2006; Li and Song, 2012) and/or reection seismic data (Taylor
and Hayes, 1983; Franke et al., 2011; Zhu et al., 2012; Li and
Song, 2012). According to Briais et al. (1993), magnetic anomaly
6 (20.45 Ma) would lie at the location of OBS-17 on the southern
limb (at km 750, see Fig. 2). Granite was dredged no more than
20 km south of picked anomaly 6. The northern counterpart of
anomaly 6 was not recognized in their model, but northern 5e (19
Ma) is identied close to OBS-19. On this same southern ank,
Barckhausen and Roeser (2004) identied anomaly 6B (23 Ma) at
the location of anomaly 6 proposed by Briais et al. (1993). Their
model predict oceanic ages that are systematically 3e5 myrs older
than that of Briais et al. (1993), the end of spreading being around
20.5 Ma instead of 15.5 Ma. Another difference is that the sepa-
ration rate jumps to 8 cm/yr in Barckhausen and Roeser (2004),
twice the rate quoted in Briais et al. (1993). There is thus a
consensus on the oceanic nature of the crust, but not on the
location of the anomalies, the age of spreading and the rate of
opening. However, in both models, true oceanic crust seems to be
present close to the continental slope, leaving very little space for a
COT (Fig. 8).
The nal velocity model shows that at both margins, velocities
higher than 6.5 km/s are remarkably at from the oceanic domain
toward the continental slope before a sudden drop downward
marked by an abrupt hinge (at 550 and 760 km, Fig. 4B). We suggest
that this velocity change marks the limit of the truly continental
crust. Together with the identication of the oldest magnetic, the
COTs appear to be slightly asymmetric: the transition width is no
larger than 60 kmon the northern side, and no more than 30 kmon
the southern side (Fig. 4B).
Only few seismic reection proles run close to the southern
margin COT and no seismic line with sufciently good overall
quality has been published at the vicinity of northern margin COT.
One seismic line published by Ding et al. (2013) crosses the
southern COT in the Dangerous Grounds area, w100 km SW and
parallel to PR2. These authors identied a detachment fault dipping
northward and rooted in a band of highly reected phases inter-
preted as the Moho. Smaller normal faults, with similar dip,
develop above the hanging wall of the detachment fault. The ac-
tivity of these faults coincides with the nal stage of rifting. The
velocity model shows localized narrow sedimentary basins at both
COTs, at 580 km and 760 km (Fig. 4B). The northern basin is more
pronounced with an average thickness of 1.5 km. The southern
basin, w1 km thick, is similar to the small basin imaged at the foot
of the slope of the southern continental margin on a nearby seismic
line (Ding et al., 2013). At the southern COT, relatively high velocity
cores seem to rise at shallow depth belowthe sedimentary basin at
760 km (corresponding to w3.5 km/s below the seaoor; Fig. 4B).
Table 2
Summary of the number of picked seismic phases, model residuals and c
2
.
Seismic phases Number of
valid data
(RMS)
residual
Average picking
uncertainty (ms)
c
2
PR1 Pg, Pn 31570 0.163 144 1.66
PmP 4359 0.126 150 0.71
Total 35929 0.159 145 1.55
PR2 Pg, Pn 19531 0.157 126 2.12
PmP 2025 0.154 150 1.05
Total 21556 0.157 128 2.02
T. Pichot et al. / Marine and Petroleum Geology xxx (2013) 1e17 12
Please cite this article in press as: Pichot, T., et al., Deep crustal structure of the conjugate margins of the SW South China Sea from wide-angle
refraction seismic data, Marine and Petroleum Geology (2013), http://dx.doi.org/10.1016/j.marpetgeo.2013.10.008
Figure 6. Plots of 1D velocity models extracted from PR1 and PR2 (see Fig. 4) in function of the depth below the sea level (A and B, respectively). C) Average velocity model of PR1
and PR2 in gray and black, respectively. Plots of the 1D velocity models of PR1 (D) and PR2 (E) in the rst few kilometers below the seaoor (sediment velocities). Plots of the 1D
velocity models of PR1 (F) and PR2 (G) in function of depth below the basement (crust and shallow upper mantle). H) Average 1D velocity model of the crust and the shallow upper
mantle from PR1 and PR2 compared to the velocity model of the average continental margin type published by Mooney et al. (1998) has been added.
Figure 7. Plot of the thickness of the whole continental crust (black), the lower crust (dark gray) and upper and middle crust (light gray). A: the northern margin and B: the southern
margin. The top of the upper crust and the top of the lower crust are determined using seismic velocity of 5.7 km/s and 6.5 km/s respectively. The Moho interface marks the base of
the crust. Here the COTs and oceanic crust are not included.
T. Pichot et al. / Marine and Petroleum Geology xxx (2013) 1e17 13
Please cite this article in press as: Pichot, T., et al., Deep crustal structure of the conjugate margins of the SW South China Sea from wide-angle
refraction seismic data, Marine and Petroleum Geology (2013), http://dx.doi.org/10.1016/j.marpetgeo.2013.10.008
The crust at the COTs is extremely thin, between 3 and 5 km on
the northern side and 3 km on the southern side (Figs. 4B and 8),
but we have no evidence for mantle exhumation. In addition to the
quite large OBS spacing, the method we use may however be
inadequate to demonstrate exhumation. Zelt et al. (2003) produced
refraction travel time tomography inversion on the Iberia margin
showing a somewhat similar velocity structures at the COT, at the
exact place where exhumation of upper mantle peridotites were
proven by ODP drilling results. The tonalite and the monzogranite
dredged close to OBS-13 and OBS-15, respectively may be inter-
preted either as the dislocation of a granitic massif (LVB A, Fig. 8)
through complex faulting during the nal stage of rifting (Fig. 8;
Reston, 2009), or evidence for exhumation of deeper level
restricted to the crust. Better resolution of the structures present at
the COTs would require forward layer modeling, with interpreta-
tion of all refracted and reected phases, using coincident reection
seismic data, as suggested by Zelt and Smith (1992).
4.4. Other refraction seismic experiments in the South China Sea
More refraction seismic proles have been published across the
northern margin (Nissen et al., 1995; Xia et al., 1998; Pin et al.,
2001; L et al., 2011; Qiu et al., 2001; McIntosh et al., 2005;
Wang et al., 2006; Wei et al., 2011; Wu et al., 2012) than across
the southern margin (Qiu et al., 2011; Ruan et al., 2011). Large
discrepancies in the design of the refraction seismic experiments
and in the modeling strategies make it difcult to compare in de-
tails one segment of rifted margin to another. Nevertheless, some
remarks regarding the rst order characteristics can be made.
Abrupt thinning of the continental crust toward the COT occurs
preferentially over the western segment of the SCS (between
Maccleseld Bank and Reed Bank area; Fig. 1) where the width of
the COT is smaller (w50 km compare to w120 km to the northeast;
Pin et al., 2001; Wei et al., 2011). Along the eastern SCS margins, the
continental crust gently thins toward the oceanic crust over an area
several hundred kilometers wide. Hayes and Nissen (2005) calcu-
late that the total amount of continental extension involved in the
rifting is nearly twice in the western part of the SCS than in the
eastern segment. These rst order comparisons indicate clear
variability of the rifting processes from east to west, in part related
to a propagating rift effect. It has long been recognized that the SCS
margins and adjacent oceanic basin display the typical V-shape
geometry of propagating rift systems (Hayes, 1985; Huchon et al.,
1998, 2001). Higher amount of stretching is reached toward the
rift tip in relation with the diachronous formation of the oceanic
crust (Courtillot, 1982). The lateral variability may also relate to the
nature of the pre-rift continental crust, the thermal regime and/or a
preexisting zone of weakness (Hayes and Nissen, 2005; Savva,
2013).
The lateral velocity variations that we document (i.e. presence of
LVBs, see Fig. 4B) have not been identied along the northeastern
segment of the SCS, except along one prole modeled by L et al.
(2011) with a different strategy (OBS 973-3, see Fig. 1 for localiza-
tion). These lateral velocity variations thus seem to be specic to
the widely extended continental crust that characterizes the
western segment of the SCS.
High Velocity Bodies (HVBs) at the COT (also described as High
Velocity Layer, HVL, or High Velocity Zone, HVZ) seem to be
restricted to the northeastern margin (Nissen et al., 1995; Pin et al.,
2001; Wang et al., 2006; Wei et al., 2011). HVBs found in the
Table 3
Compilation of the seismic velocity range in the sedimentary cover from previous
refraction study. See Figure 1 for localization.
Velocity range (km/s) References Locations
1.6e2.9 [uncompacted] Wang et al., 2006 OBS 2001
3e4.5 [compacted]
1.55e5.5 Nissen et al., 1995 ESP-E1985
1.55e4.875 Nissen et al., 1995 ESP-C1985
1.7e3.5 [uncompacted] Pin et al., 2001 OBS 1993
3.5e5.5 [compacted]
1.9e3.6 Wu et al., 2012 OBS 2006-1
1.55e5.25 Nissen et al., 1995 ESP-W1985
1.7e4.5 Qiu et al., 2001 OBS 1996-IV
2.1e4.5 L et al., 2011 OBS 973-3
2.5e4.5 Qiu et al., 2011 OBS 973-1
Figure 8. Geological interpretation of the crustal structure of the SW SCS conjugate margins based on our seismic refraction prole and others observations provided by seismic
reection data in the vicinity of the refraction prole. Seismic velocities are used to delineate the uncompacted sediment (white), the consolidated sediments (very light gray), the
upper and middle crust (light gray) and the lower crust (dark gray). Because of the poor resolution of the velocity model in the oceanic crust, all the oceanic crust is represented in
the very dark gray. Shallowsmall-scale normal faulting and minor large-scale normal faults are shown in thin black lines. Major large-scale normal faults are in bold black lines. Low
Velocity Bodies (LVBs) are indicated by small number along the northern margin and by letters along the southern margin. Dredged samples have been projected along the OBS
prole as shown by colored stars (see Fig. 1). The oldest magnetic anomalies identied in the oceanic crust are indicated by vertical black dashed line (after Briais et al., 1993;
Barckhausen and Roser, 2004).
T. Pichot et al. / Marine and Petroleum Geology xxx (2013) 1e17 14
Please cite this article in press as: Pichot, T., et al., Deep crustal structure of the conjugate margins of the SW South China Sea from wide-angle
refraction seismic data, Marine and Petroleum Geology (2013), http://dx.doi.org/10.1016/j.marpetgeo.2013.10.008
northeastern part of the SCS spread several hundred kilometers
across the margin. A maximum thickness of 17 km was estimated
according to ESP survey (Nissen et al., 1995) and 10 km using OBS
data (Pin et al., 2001). The associated P-wave velocity ranges be-
tween 7.0 and 7.5 km/s. These HVBs have been interpreted either as
residual mac rocks inherited from the pre-rifted crustal structure
(Nissen et al., 1995) or post-rifting underplating in relation with the
recent volcanism distributed over the SCS (Pin et al., 2001; Wang
et al., 2006) or potential serpentinization of the upper mantle
(Franke et al., 2011).
Although no HVB was reported so far over the western segment
of the SCS, our nal model shows high velocity material (7.0e
7.7 km/s) at the base of the lower crust, but distributed into thin
lenses (2e3 km thick) rather than voluminous bodies. Since
underplating material is generally found with thicknesses in excess
of 5e6 km, thin lenses would tend to be interpreted as serpenti-
nized upper mantle (Watremez et al., 2011), as frequently found at
necking zones and COTs. Serpentinization is generally interpreted
as the result of the interaction of mantle peridotites with seawater
through a network of faults that link the shallow upper mantle to
the seaoor. The presence of serpentinized upper mantle has been
extensively studied at the Iberian margin (Boillot and Winterer,
1988), Newfoundland (Tucholke et al., 2007) and South Australia
(Nicholls et al., 1981). Many of our HVLs are found at the COTs and
may thus be related to serpentinization. However, a serpentinized
mantle origin for the HVL found where the crust is thicker
(Zhongjannan Basin, NE extent of the Phu Khan Basin, and between
LVBs A and B) may not be applicable. Water would have to percolate
through a w16 km thick continental crust, but since the large-scale
normal faults root in the ductile lower crust, inltration may not
reach the shallow upper mantle.
Whatever the origin of the high velocity bodies and lenses, an
Andean-type arc associated with the northward subduction of the
Pacic plate would likely produce petrological heterogeneities
within the lower continental crust, and it would not be surprising
to nd these bodies as stretched relics.
4.5. Mechanisms of extension of the SW part of the SCS
As mentioned in a previous section, the SW conjugate margins
of the SCS showmore stretching than the NE ones. The 32 km thick
continental crust beneath Hainan Island (Chen et al., 2010) is
generally considered as the original unstretched crust. Off Hainan,
the margin thins over a distance of 790 km along the northern
margin and 440 km along the southern margin (ending with Pal-
awan Trough), which makes it one of the widest rifted margins
worldwide, with a total width of extended continental crust of
1230 km. It is comparable in size with the 900 km-long Basin and
Range province in the western United States (Parsons, 1995).
Extension across the conjugate margins of the SW SCS is
distributed on small-scale (15e30 km) and largeescale (45e90 km)
normal faulting described in Sections 4.1 and 4.2. Following the
study of Zuber et al. (1986) in the Basin and Range, these two
wavelengths of extensional deformation may directly relate to the
presence of competent layers, here, the upper and middle crust and
the shallow upper mantle, separated by a ductile lower crust. Here,
the stretching is distributed symmetrically over both conjugate
margins. It has been mentioned in Section 3.2 that the spacing of
the LVBs decreases as the continental crust thins along the northern
margin of the SW sub-basin of the SCS. Comparable decrease was
described at the Iberian margin (Prez-Gussiny et al., 2003;
Ranero and Prez-Gussiny, 2010; Sutra and Manatschal, 2012).
We estimate the total amount of extension along the SW con-
jugate margins of the SCS by assuming conservation of the volume
of continental crust. The total surface of continental crust along the
modeled section is estimated using the top of the crust, dened by
the 5.7 km/s isocontour velocity and the base of the crust, dened
by the Moho interface. This results in a surface (or volume per unit
margin width) of the fully continental crust (the COT areas are not
included) of approximately 11,000 km
2
. Eight main large offset
normal faults have been interpreted from the nal velocity model
(Fig. 8). Considering a thickness of 32 km as the initial thickness of
the pre-rift continental crust, each normal fault zone would have to
accommodate 60 km of horizontal extension, which is higher than
the most optimistic value (20 km) assigned for such large-scale
normal fault described by Hayes et al. (1995). This stretching
discrepancy has been widely discussed for the SCS (Clift and Lin,
2001) and at many other margins, and 3 mechanisms are now
proposed to resolve it: depth-dependent stretching (Davis and
Kusznir, 2004; Lavier and Manatschal, 2006; Huismans and
Beaumont, 2011), polyphase faulting (Reston, 2005; Prez-
Gussiny et al., 2003), sequential faulting (Ranero and Prez-
Gussiny, 2010). The rst mechanism would require a region of
highly stretched upper crust (to compensate for the region of highly
thinned lower crust and keep both crusts in contact at the end,
Reston, 2005, 2009), or out-section ow or exhumation of the
lower crust (toward the oceanic domain, toward the unextended
margins, toward regions outside the prole). The volume of lower
crust involved would be so large that this solution, wherever the
lower crust would go, seems unrealistic. On the other hand, the
resolution of the seismic imaging generally underestimates the
amount of extension measurable from the observed faulting.
Recent studies (Reston, 2005; Ranero and Prez-Gussiny, 2010)
have shown that the extension discrepancy can be resolved by
accurately measure fault extension and compare it with crustal
thinning, using depth-migrated seismic reection proles.
Following these results, we favor solutions that do not require high
differential stretching, pervasive extension in the upper crust
accompanying large-scale normal faulting.
5. Conclusions
1) The continental crusts of the conjugate margins of the SWSouth
China Sea sub-basin show similar large scale seismic velocities
and structural (faults and fabric) characteristics and are thus
genetically linked. One large granitic massif with low seismic
velocity bounds the southern margin at the edge of the oceanic
crust. Based on dredged material (monzogranite), this massif
belongs to the Yanshanian plutons well studied in North and
South China and points to a compositional and structural het-
erogeneity of the continental crust prior to stretching.
2) Most of the continental crust deforms in a brittle manner. Two
wavelengths of velocity perturbations are imaged that correlate
well with faults observed on seismic reection proles. Small-
scale normal faults (grabens, horsts and rotated faults blocks)
are often associated with tilt of the velocity isocontours within
the upper crust at a 15e30 km wavelength. The mid-crust at a
depth of 5e15 km below seaoor shows the largest lateral ve-
locity variation, thus dening low velocity bodies bounded by
large-scale normal faults rooted in the lower crust. A larger than
expected number of these large-scale normal faults seem to be
present along the northern margin. These faults show no pref-
erential vergence. Major sedimentary basins are located above
the low velocity bodies. Along the northern margin, spacing
decreases from 90 to 45 km as the total crust thins toward the
ContinenteOcean Transition.
3) The lower crust has been deforming in a ductile regime. The
Moho interface remains rather at over the large-scale normal
faults bounding lowvelocity bodies, despite a great variability in
the size of the sedimentary basins, suggesting that isostatic
T. Pichot et al. / Marine and Petroleum Geology xxx (2013) 1e17 15
Please cite this article in press as: Pichot, T., et al., Deep crustal structure of the conjugate margins of the SW South China Sea from wide-angle
refraction seismic data, Marine and Petroleum Geology (2013), http://dx.doi.org/10.1016/j.marpetgeo.2013.10.008
compensation is achieved through the ductile lower crust. The
main large-scale normal faults root within this level.
4) The ContinenteOcean Transitions are narrow and slightly
asymmetric: 60 km on the northern side, and no more than
30 km on the southern side. High velocity material is present at
very shallow level at these transitions (3 km below the sedi-
mentary basin), but exhumation of upper mantle peridotites or
serpentinites cannot be proven.
5) Large-scale extensional structures observed along the conjugate
margins of the SW South China Sea fail to explain the total
amount of horizontal extension estimated from thinning. To
explain the discrepancy, pervasive faulting/deformation not
detected on seismic proles is needed.
References
Barckhausen, U., Roeser, H.A., 2004. Seaoor spreading anomalies in the South
China Sea revisited. In: Clift, P., Kuhnt, W., Wang, P., Hayes, D.E. (Eds.), Conti-
nent-ocean Interactions in the East Asian Marginal Seas. Geophys. Monograph
Series. American Geophysical Union, pp. 121e125.
Boillot, G., Winterer, E., 1988. Drilling on the Galicia margin: retrospect and pros-
pect. In: Proceedings Ocean Drilling Program. Scientic Results, vol. 103,
pp. 809e828.
Braitenberg, C., Wienecke, S., Wang, Y., 2006. Basement structures from satellite-
derived gravity eld: south China Sea ridge. J. Geophys. Res. 111, 1e15. http://
dx.doi.org/10.1029/2005JB003938.
Briais, A., Patriat, P., Tapponnier, P., 1993. Updated interpretation of magnetic
anomalies and sea oor spreading stages in the South China Sea: implications
for the Tertiary tectonics of Southeast Asia. J. Geophys. Res. 98, 6299e6328.
Chamot-Rooke, N., Gaulier, J.-M., Jestin, F., 1997. Constraints on Moho depth and
crustal thickness in the Liguro-Provenal basin from a 3D gravity inversion:
geodynamic implications. Revue de lInstitut Franais du Ptrole 52, 557e583.
Chen, Y., Niu, F., Liu, R., Huang, Z., Tkalcic, H., Sun, L., Chan, W., 2010. Crustal
structure beneath China from receiver function analysis. J. Geophys. Res. 115
(B03307), 1e22. http://dx.doi.org/10.1029/2009JB006386.
Christensen, N., Mooney, W., 1995. Seismic velocity structure and composition of
the continental crust: a global view. J. Geophys. Res. 100, 9761e9788.
Clift, P.D., Lin, J., 2001. Preferential mantle lithospheric extension under the South
China margin. Mar. Pet. Geology. 18, 929e945.
Clift, P.D., Lin, J., Barckhausen, U., 2002. Evidence of low exural rigidity and low
viscosity lower continental crust during continental break-up in the South
China Sea. Mar. Pet. Geology. 19, 951e970.
Courtillot, V., 1982. Propagating rift and continental break up. Tectonics 1 (3), 239e250.
Davis, M., Kusznir, N., 2004. Depth-Dependent lithospheric stretching at rifted
continental margins. In: Karner, G.D. (Ed.), Proceedings of NSF Rifted Margins
Theoretical Institute. Columbia University Press, pp. 92e136.
Ding, W., Franke, D., Li, J., Steuer, S., 2013. Seismic stratigraphy and tectonic struc-
ture from a composite multi-channel seismic prole across the entire
dangerous grounds, South China Sea. Tectonophysics 582, 162e176.
Franke, D., 2012. Rifting, lithosphere breakup and volcanism: comparison of
magma-poor and volcanic rifted margins. Mar. Pet. Geol., 1e25. http://
dx.doi.org/10.1016/j.marpetgeo.2012.11.003.
Franke, D., Barckhausen, U., Baristeas, N., Engels, M., Lutz, R., Montano, J., 2011. The
continent-ocean transition at the southeastern margin of the South China Sea.
Mar. Pet. Geol. 28, 1187e1204.
Gilder, S.A., Keller, G.R., Luo, M., Goodell, P.C., 1991. Timing and spatial distribution
of rifting in China. Tectonophysics 197, 225e243.
Hayes, D.E., 1985. Margins of the southwest sub-basin of the South China Sea-A
frontier exploration target? Energy 10 (3/4), 373e382.
Hayes, D.E., Nissen, S.S., 2005. The south China Sea margins: implication for rifting
contrast. Earth Planet. Sci. Lett. 237, 601e616.
Hayes, D.E., Nissen, S.S., Buhl, P., Diebold, J., Yao, B., Zeng, W., Chen, Y., 1995.
Throughgoing crustal faults along the northern margin of the South China Sea
and their role in crustal extension. J. Geophys. Res. 100, 22435e22446.
Huchon, P., Nguyen, T., Chamot-Rooke, N., 1998. Finite extension across the South
Vietnam basins from 3D gravimetric modelling: relation to South China Sea
kinematics. Mar. Pet. Geology. 15, 619e634.
Huchon, P., Nguyen, T., Chamot-Rooke, N., 2001. Propagation of continental breakup
in the southwestern South China Sea. In: Wilson, R., Whitmarsh, R., Taylor, B.,
Froitzheim, N. (Eds.), Geological Society of London, vol. 187, pp. 31e50.
Huismans, R., Beaumont, C., 2003. Symmetric and asymmetric lithospheric exten-
sion: relative effects of frictional-plastic and viscous strain softening.
J. Geophys. Res. 108 (B10), 2496. http://dx.doi.org/10.1029/2002JB002026.
Huismans, R., Beaumont, C., 2011. Depth-dependent extension, two-stage breakup
and cratonic underplating at rifted margins. Nature 473, 74e79.
Huismans, R.S., Buiter, S.J.H., Beaumont, C., 2005. The effect of plastic-viscous
layering and strain-softening on mode selection during lithospheric exten-
sion. J. Geophys. Res. 110 (B02406). http://dx.doi.org/10.1029/2004JB003114.
Hutchison, C.S., Vijayan, V.R., 2010. What are the Spratly islands? J. Asian Earth Sci.
39 (5), 371e385.
Jahn, B.M., Chen, P.Y., Yen, T.P., 1976. Rb-Sr ages of granitic rocks in southeastern
China and their tectonic signicance. Geol. Soc. America Bull. 87, 763e776.
Korenaga, J., 2003. Tomo2D: A C Package for 2-D Joint Refraction and Reection
Travel-time Tomography. http://people.earth.yale.edu/prole/jk525/content/
software.
Korenaga, J., Kelemen, P.B., 1997. Origin of gabbro sills in the Moho transition zone
of the Oman ophiolite: implications for magma transport in the oceanic lower
crust. J. Geophys. Res. 102, 27729e27749.
Korenaga, J., Holbrook, W.S., Kent, G.M., Kelemen, P.B., Detrick, R.S., Larsen, H.,
Hopper, J.R., Carlo, M., 2000. Crustal structure of the southeast Greenland
margin from joint refraction and reection seismic tomography. J. Geophys. Res.
105, 591e614.
Kudrass, H.W., Wiedicke, M., Cepek, P., Kreuzer, H., M
Q
ller, P., 1986. Mesozoic and
Cainozoic rocks dredged from the South China Sea (Reed Bank area) and Sulu
Sea and their signicance for plate-tectonic reconstructions. Mar. Pet. Geology.
3, 19e30.
Lavier, L.L., Manatschal, G., 2006. A mechanism to thin the continental lithosphere
at magma-poor margins. Nature 440, 324e328.
Li, C.F., Song, T.R., 2012. Magnetic recording of the Cenozoic oceanic crustal accre-
tion and evolution of the South China Sea basin. Chin. Sci. Bull. 57 (24), 3165e
3181. http://dx.doi.org/10.1007/s11434-012-5063-9.
L, C.C., Hao, T., Qiu, X.L., Zhao, M.H., You, Q.Y., 2011. A study on the deep structure
of the northern part of southwest sub-basin from ocean bottom seismic data,
South China Sea. Chin. J. Geophys. 54 (12), 3129e3138 (in Chinese).
McIntosh, K.D., Nakamura, Y., Wang, T.K., Shih, R.C., Chen, A.T., Liu, C.S., 2005.
Crustal-scale seismic proles across Taiwan and the western Philippine Sea.
Tectonophysics 401, 23e54.
McKenzie, D., 1978. Some remarks on the development of sedimentary basins. Earth
Planet. Sci. Lett. 40, 25e42.
Mooney, W., Laske, G., Masters, T., 1998. CRUST 5.1: a global crustal model at 5N-5.
J. Geophys. Res. 103, 727e747.
Nguyen, N.T., Nguyen, T.T.H., 2013. Topography of the Moho and Earth crust
structure beneath the east Vietnam sea from 3D inversion of gravity eld data.
Acta Geophys. 61 (2), 357e384. http://dx.doi.org/10.2478/s11600-012-0078-9.
Nicholls, I.A., Ferguson, J., Jones, H., Marks, G.P., Mutter, J.C., 1981. Ultramac blocks
from the seaoor southwest of Australia. Earth Planet. Sci. Lett. 56, 362e374.
Nissen, S.S., Hayes, D.E., Buhl, P., Diebold, J., Bochu, Y., Weijun, Z., Yongqin, C., 1995.
Deep penetration seismic sounding across the nothern margin of the South
China Sea. J. Geophys. Res. 100, 22407e22433.
Parker, R.L., 1973. The rapid calculation of potential anomalies. Geophys. J. R. As-
tronomical Soc. 31 (4), 447e455. http://dx.doi.org/10.1111/j.1365-
246X.1973.tb06513.x.
Parsons, T., 1995. The basin and range province. In: Olsen, K. (Ed.), Continental Rifts:
Evolution, Structure and Tectonics. Elsevier, Amsterdam, ISBN 044489-566-3,
pp. 277e324.
Prez-Gussiny, M., Ranero, C.R., Reston, T.J., Sawyer, D., 2003. Structure and
mechanisms of extension at the Galicia Interior Basin, west of Iberia. J. Geophys.
Res. 108 (B5), 2245. http://dx.doi.org/10.1029/2001JB000901.
Pin, Y., Di, Z., Zhaoshu, L., 2001. A crustal structure across the northern continental
margin of the South China Sea. Tectonophysics 338, 1e21.
Pubellier, M., Ego, F., Chamot-Rooke, N., Rangin, C., 2003. The building of pericra-
tonic mountain ranges: structural and kinematic constraints applied to GIS-
based reconstructions of SE Asia. Bull. de la Socit gologique de France 174
(6), 561e584.
Qiu, X.L., Ye, S., Wu, S., Shi, X., Zhou, D., Xia, K., Flueh, E.R., 2001. Crustal structure
across the Xisha trough, northwestern South China Sea. Tectonophysics 341,
179e193.
Qiu, X.L., Zhao, M.H., Ao, W., L, C.C., Hao, T.Y., You, Q.Y., Ruan, A.G., Li, J.B., 2011. OBS
survey and crustal structure of the SW sub-basin and Nansha block, South
China Sea. Chin. J. Geophys. 54 (12), 1009e1021.
Ranero, C.R., Prez-Gussiny, M., 2010. Sequential faulting explains the asymmetry
and extension discrepancy of conjugate margins. Nature 468, 294e300. http://
dx.doi.org/10.1038/nature09520.
Reston, T.J., 2005. Polyphase faulting during the development of the west Galicia
rifted margin. Earth Planet. Sci. Lett. 237, 561e576.
Reston, T.J., 2009. The structure, evolution and symmetry of the magma-poor
rifted margins of the North and Central Atlantic: a Synthesis. Tectonophysics
468, 6e27.
Ruan, A. G., Niu, X. W., Qiu, X. L., Li, J. B., Wu, Z. L., Zhao, M. H., Wei, X. D., 2011. (in
Chinese) 54(12), 3139e3149.
Rudnick, R.L., Fountain, D.M., 1995. Nature and composition of the continental
crust: a lower crustal perspective. Rev. Geophys. 33, 267e309.
Sandwell, D.T., Smith, W.H.F., 2009. Global marine gravity from retracked Geosat
and ERS-1 altimetry: ridge segmentation versus spreading rate. J. Geophys. Res.
114 (B01411), 1e18. http://dx.doi.org/10.1029/2008JB006008.
Savva, D., 2013. Variabilit des processus dextension continentale en mer de chine
du Sud. Universit Pierre et Marie Curie, Paris. Ph.D. Thesis.
Schluter, H.U., Hinz, K., Block, M., 1996. Tectono-stratigraphic terranes and
detachment faulting of the South China Sea and Sulu sea. Mar. Geology. 130,
39e51.
Smith, W.H.F., Sandwell, D.T., 1997. Global seaoor topography from satellite
altimetry and ship depth soundings. Science 277, 1957e1962.
T. Pichot et al. / Marine and Petroleum Geology xxx (2013) 1e17 16
Please cite this article in press as: Pichot, T., et al., Deep crustal structure of the conjugate margins of the SW South China Sea from wide-angle
refraction seismic data, Marine and Petroleum Geology (2013), http://dx.doi.org/10.1016/j.marpetgeo.2013.10.008
Sutra, E., Manatschal, G., 2012. Hoes does the continental crust thin in a hyperextended
rifted margin? Insights from the Iberian margin. Geology 40 (2), 139e142.
Taylor, B., Hayes, D.E., 1983. Origin and history of the South China Sea basin. In:
Hayes, D.E. (Ed.), The Tectonic and Geologic Evolution of Southeast Asian Sea
and Islands. AGU, Washington, DC, pp. 23e56.
Toomey, D.R., Solomon, S.C., Purdy, G.M., 1994. Tomographic imaging of the shallow
crustal structure of the East Pacic Rise at 9N30N. J. Geophys. Res. 99, 24135e
24157.
Tucholke, B.E., Sawyer, D.S., Sibuet, J.-C., 2007. Breakup of the NewfoundlandeIberia
rift. In: Geological Society of London, Special Publication, vol. 282, pp. 9e46.
VanAvendonk, H.J.A., Harding, A.J., Orcutt, J.A., 1998. Atwo-dimensional tomographic
study of the Clipperton transform fault. J. Geophys. Res. 103, 17885e17899.
Wang, T.K., Chen, M.K., Lee, C.S., Xia, K., 2006. Seismic imaging of the transitional
crust across the northeastern margin of the South China Sea. Tectonophysics
412, 237e254.
Wang, Y., Fan, W., Zhang, G., Zhang, Y., 2012. Phanerozoic tectonics of the South
China Block: Key observations and controversies. Gondwana Res.. http://
dx.doi.org/10.1016/j.gr.2012.02.019.
Watremez, L., Leroy, S., Rouzo, E., DAcremont, E., Unternehr, P., Ebinger, C.,
Lucazeau, F., Al-Lazki, A., 2011. The crustal structure of the north-eastern Gulf of
Aden continental margin: insights from wide-angle seismic data. Geophys. J.
Int. 184, 575e594. http://dx.doi.org/10.1111/j.1365-246X.2010.04881.x.
Wei, X.D., Ruan, A.G., Zhao, M.H., Qiu, X.L., Li, J.B., Zhu, J.J., Wu, Z.L., Ding, W.W., 2011.
A wide-angle OBS prole across the Dongsha uplift and Chaoshan depression in
the mid-northern South China Sea. Chin. J. Geophys. 54 (12), 1149e1160.
Wernicke, B., 1985. Uniform-sense normal simple shear of the continental litho-
sphere. Can. J. Earth Sci. 22, 108e125.
Wu, Z., Li, J., Ruan, A., Lou, H., Ding, W.W., Niu, X., Li, X.B., 2012. Crustal structure of
the northwestern subbasin, South China Sea: Results from a wide-angle seismic
experiment. Earth Sci. Sci. China 55 (1), 159e172.
Xia, K., Xia, K., Zhou, D., Su, D., Flueh, E.R., Ye, S., He, H., Chen, W., Xie, Y.,
Wang, G., Liu, S., Fu, G., Wang, J., Chen, J., 1998. The velocity structure of the
Yinggehai Basin and its hydrocarbon implication. Chin. Sci. Bull. 43, 2047e
2054.
Yan, Q., Shi, X., Liu, J., Wang, K., Bu, W., 2010. Petrology and geochemistry of
Mesozoic granitic rocks from the Nansha micro-block, the South China Sea:
constraints on the basement nature. J. Asian Earth Sci. 37, 130e139.
Zelt, C.A., Barton, P.J., 1998. Three-dimensional seismic refraction tomography: a
comparison of two methods applied to date from the Faeroe Basin. J. Geophys.
Res. 103, 7187e7210.
Zelt, C.A., Forsyth, D., 1994. Modeling wide-angle seismic data for crustal structure:
southeastern Grenville Province. J. Geophys. Res. 99 (B6), 11687e11704.
Zelt, C.A., Smith, R.B., 1992. Seismic traveltime inversion for 2-D crustal velocity
structure. Geophys. J. Int. 108, 16e34.
Zelt, C.A., Sain, K., Naumenko, J.V., Sawyer, D.S., 2003. Assessment of crustal velocity
models using seismic refraction and reection tomography. Geophys. J. Int. 153,
609e626.
Zhang, Z.J., Liu, Y.F., Zhang, S.F., Zhang, G.C., Fan, W.M., 2009. Crustal P-wave velocity
structure and layering beneath ZhujiangkoueQiongdongnan basins in the
northern continental margin of South China Sea. Chin. J. Geophys. 52 (5), 1012e
1023.
Zhu, J., Qiu, X., Kopp, H., Xu, H., Sun, Z., Ruan, A., Sun, J., Wei, X., 2012. Shallow
anatomy of a continenteocean transition zone in the northern South China Sea
from multichannel seismic data. Tectonophysics 554e557, 18e29.
Ziegler, P.A., 2001. Dynamic Processes Controlling Development of Rifted Basins.
EUCOR-URGENT Publication 1, pp. 1e51.
Zuber, M.T., Parmentier, E.M., Fletcher, R.C., 1986. Extension of continental litho-
sphere: a model for two scales of basin and range deformation. J. Geophys. Res.
91, 4826e4838.
T. Pichot et al. / Marine and Petroleum Geology xxx (2013) 1e17 17
Please cite this article in press as: Pichot, T., et al., Deep crustal structure of the conjugate margins of the SW South China Sea from wide-angle
refraction seismic data, Marine and Petroleum Geology (2013), http://dx.doi.org/10.1016/j.marpetgeo.2013.10.008

Vous aimerez peut-être aussi