ArticlePDF Available

Abstract and Figures

The Dalrymple Trough is a 150-km-long, 30-km-wide basin located at the northern termination of the Owen Fracture Zone (OFZ), which is the present-day active India-Arabia plate boundary. The Dalrymple Trough is closely associated with the Murray Ridge, a complex of prominent bathymetric highs located on its eastern flank. Recent multibeam mapping of the connection between the Dalrymple Trough and the OFZ revealed a horsetail structure, which suggests a close relationship between geological histories of both structures. However, the 3-6 Ma age of initiation of the OFZ contrasts with the commonly accepted Early Miocene emplacement of the Dalrymple Trough. Recent seismic lines document a new tectonic history of the Dalrymple Trough, involving two major episodes of deformation along the India-Arabia plate boundary at ~8-10 Ma and ~1.9 ± 0.9 Ma. The 8-10 Ma episode is marked by a system of folds linked to the main uplift of the southern Murray Ridge and the first uplift of the northern Murray Ridge. This episode is related to a global plate reorganization event in the Late Miocene, well expressed by intraplate deformation in the Central Indian Ocean. The Dalrymple Trough opened at ~1.9 ±0.9 Ma subsequently to the formation of a stepover at the India-Arabia plate boundary, coeval with the regional M-unconformity in the Oman abyssal plain, which marks a structural reorganization of the Makran accretionary wedge, and the last uplift of the northern Murray Ridge.
No caption available
… 
No caption available
… 
Content may be subject to copyright.
Tectonics of the Dalrymple Trough and uplift of the Murray Ridge (NW
Indian Ocean)
Mathieu Rodriguez, Nicolas Chamot-Rooke, Philippe Huchon, Marc
Fournier, Siegfried Lallemant, Matthias Delescluse, S´ebastien Zaragosi,
Nicolas Mouchot
PII: S0040-1951(14)00450-8
DOI: doi: 10.1016/j.tecto.2014.08.001
Reference: TECTO 126408
To appear in: Tectonophysics
Received date: 2 April 2014
Revised date: 1 August 2014
Accepted date: 2 August 2014
Please cite this article as: Rodriguez, Mathieu, Chamot-Rooke, Nicolas, Huchon,
Philippe, Fournier, Marc, Lallemant, Siegfried, Delescluse, Matthias, Zaragosi, S´ebastien,
Mouchot, Nicolas, Tectonics of the Dalrymple Trough and uplift of the Murray Ridge
(NW Indian Ocean), Tectonophysics (2014), doi: 10.1016/j.tecto.2014.08.001
This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
1
Tectonics of the Dalrymple Trough and uplift of the Murray Ridge (NW Indian Ocean)
Mathieu Rodriguez
1*
, Nicolas Chamot-Rooke
1
, Philippe Huchon
2,3
, Marc Fournier
2,3
, Siegfried
Lallemant
4
, Matthias Delescluse
1
, Sébastien Zaragosi
5
, Nicolas Mouchot
6
(1) Laboratoire de Géologie de l'Ecole Normale Supérieure, CNRS UMR 8538, 24 rue Lhomond,
75005 Paris, France
(2) Institut des Sciences de la Terre de Paris, UMR 7193, Université Pierre & Marie Curie, case 129, 4
place Jussieu, 75005 Paris, France
(3) iSTeP, UMR 7193, CNRS, F-75005 Paris, France
(4) Département Géosciences Environnement, Université de Cergy-Pontoise, 5 mail Gay-Lussac,
Neuville/Oise, 95031 Cergy-Pontoise, France
(5) EPOC Université de Bordeaux, UMR 5805, avenue des facultés, 33405 Talence, France
(6) Beicip-Franlab, 232 Avenue Napoléon Bonaparte, 92500 Rueil-Malmaison, France
*Corresponding author: rodriguez@geologie.ens.fr
ABSTRACT
The Dalrymple Trough is a 150-km-long, 30-km-wide basin located at the northern termination
of the Owen Fracture Zone (OFZ), which is the present-day active India-Arabia plate boundary.
The Dalrymple Trough is closely associated with the Murray Ridge, a complex of prominent
bathymetric highs located on its eastern flank. Recent multibeam mapping of the connection
between the Dalrymple Trough and the OFZ revealed a horsetail structure, which suggests a
close relationship between geological histories of both structures. However, the 3-6 Ma age of
initiation of the OFZ contrasts with the commonly accepted Early Miocene emplacement of the
Dalrymple Trough. Recent seismic lines document a new tectonic history of the Dalrymple
Trough, involving two major episodes of deformation along the India-Arabia plate boundary at
~8-10 Ma and ~1.9 ± 0.9 Ma. The 8-10 Ma episode is marked by a system of folds linked to the
main uplift of the southern Murray Ridge and the first uplift of the northern Murray Ridge.
This episode is related to a global plate reorganization event in the Late Miocene, well expressed
by intraplate deformation in the Central Indian Ocean. The Dalrymple Trough opened at ~1.9 ±
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
2
0.9 Ma subsequently to the formation of a stepover at the India-Arabia plate boundary, coeval
with the regional M-unconformity in the Oman abyssal plain, which marks a structural
reorganization of the Makran accretionary wedge, and the last uplift of the northern Murray
Ridge.
1. Introduction
The Dalrymple Trough consists of a series of basins forming the present-day loose boundary between
India and Arabia plates in the Arabian Sea, connecting the pure dextral strike-slip Owen Fracture Zone
(OFZ hereafter) to the sinistral Ornach-Nal Fault Zone in Pakistan (Fig. 1; McKenzie and Sclater,
1971; Minshull et al., 1992). The trough is located south of the Makran subduction zone, which
absorbs the convergence between the Arabian and Eurasian plates and has produced strong
earthquakes in the recent past (M
w
=8.1 for the 1945 event). The trough is flanked to the east by the
Murray Ridge complex, a series of bathymetric highs standing off India and Pakistan.
Seismicity along the entire India-Arabia boundary is scarce, but the strongest magnitude event
(M
w
=5.8) has been recorded in the Dalrymple Trough (Fig. 2) (Quittmeyer and Kafka, 1984; Gordon
and DeMets, 1989; Fournier et al., 2001). Kinematic models suggest that the present-day opening rate
of the Dalrymple Trough does not exceed a few millimeters per year (DeMets et al., 2010; Fournier et
al., 2011), with various amounts of transtension.
Multibeam mapping revealed a fault pattern at the southern part of the Dalrymple Trough typical of a
horsetail termination (Fig. 1, 2) (Fournier et al., 2011; Rodriguez et al., 2011). Few active horsetail
terminations have been documented so far in deep-sea environments. These include the northern
Andaman Sea at the termination of the Sagaing fault in SE Asia (Pubellier et al., 2005; Cattin et al.,
2009; Morley et al., 2013), and the North Aegean Trough at the termination of the North Anatolian
fault in the Aegean Sea (Laigle et al., 2000; Papanikolaou et al., 2002; McNeill et al., 2004).
Although the structure of the Dalrymple Trough has been well characterized by previous seismic
studies (Edwards et al., 2000, 2008; Gaedicke et al., 2002a,b), the way it connects to the OFZ and how
it relates to its tectonic history remain unknown. The main misfit concerns the Pliocene age of the
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
3
OFZ (Fournier et al., 2008a,b; 2011; Rodriguez et al., 2011) and the inferred Early Miocene age of the
Dalrymple Trough-Murray Ridge system (Gaedicke et al., 2002a,b; Clift et al., 2001).
Here we present a set of recent seismic reflection data (OWEN-2 cruise) that documents a new
stratigraphic framework for the tectonic history of the Dalrymple Trough and the Murray Ridge in
close association with structural reorganizations of the India-Arabia plate boundary at ~8-10 Ma and
~1.9 ± 0.9 Ma. The origin of the M-unconformity in the Oman abyssal plain is revised in the light of
this new tectonic framework. The first objective of this study is to understand how rifting took place in
the Dalrymple Trough, and subsequently evolved into a large and complex stepover basin. The second
objective is to unravel the complex, poly-phased history of the Murray Ridge uplift, and identify
geodynamic changes that controlled it.
2. Geological background
2.1. Morphology and structure of the Dalrymple Trough and the Murray Ridge
The Dalrymple Trough is divided into two main segments (Edwards et al., 2000). The southern part of
the Dalrymple Trough (between 22-23°N) is a 150-km-long, 30-km-wide, 4200-m-deep basin, flanked
on its eastern side by the ~400-m deep southern Murray Ridge (Fig. 1, 2). The southern part of the
Dalrymple Trough abruptly ends at the Jinnah High at ~23°N (Burgath et al., 2002), which is formed
by the tilt of the Indus deposits of the Oman abyssal plain (Ellouz-Zimmerman et al., 2007a,b). The
Jinnah High marks the transition towards the northern part of the Dalrymple Trough, which is a ~120
km-long, ~40 km-wide stepover basin. The northern Murray Ridge (between 23°N and 23°40'N)
forms a normal faulted horst with a more subtle topographic expression than the southern segment
(about 1000-m high with respect to the surrounding seafloor, Fig. 1).
The pure strike-slip or transtensive character of a stepover basin depends on the colinearity between
the direction of relative motion and that of the strike slip fault (Wu et al., 2009). Because the
Dalrymple Trough deviates from the small circle defined by Fournier et al. (2011) (Fig. 2), the
structure is considered as transtensive, in agreement with focal mechanisms of earthquakes.
The basement of the Murray Ridge, although never sampled, is interpreted as continental in origin
according to seismic refraction data (Edwards et al., 2008). The Dalrymple Trough represents a very
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
4
narrow ocean-continent transition, with abrupt crustal thickness variations from a ~14 km-thick
continental crust beneath the Murray Ridge to a 6 km-thick oceanic crust in the Oman abyssal plain
(Edwards et al., 2008). A set of volcanic reliefs, namely the Qalhat Seamount and the Little Murray
Ridge, are located west of the trough (Fig. 1, 2) (Edwards et al., 2000; Fournier et al., 2011).
2.2. Geological history
Kinematic and structural studies show that the present-day active OFZ is no older than 3 to 6 Ma
(Fournier et al., 2008; 2011; Rodriguez et al., 2011; 2013b). The exact location of the fossil India-
Arabia plate boundary remains debated (Whitmarsh, 1979; Mountain and Prell, 1990; Edwards et al.,
2000; Royer et al., 2002). The recent identification of a fracture zone buried under the Indus deposits
5-10 km east of the OFZ (Rodriguez et al., 2014), together with magnetic anomalies reconstructions
(Chamot-Rooke et al., 2009), suggest that the plate boundary remained close to the Owen Ridge since
at least the beginning of oceanic accretion in the Gulf of Aden in the Early Miocene (Fournier et al.,
2010). Buried, abrupt and sharp vertical fault offsets recognized on the eastern side of the Southern
Murray Ridge (Edwards et al., 2000) could correspond to the northward prolongation of the fracture
zone that used to form the Miocene India-Arabia plate boundary (Fig. 3).
As attested by Paleogene hemipelagites recognized on their top (Shipboard Scientific Party, 1989;
Gaedicke et al., 2002a,b), the Owen and Murray Ridges were part of a series of bathymetric highs
formed in Paleocene-Early Eocene times (Fig.1, 3) and subsequently rejuvenated during Neogene
times. Previous works (Gaedicke et al., 2002b) proposed two phases of uplift of the Murray Ridge
related to two coeval phases of subsidence in the Dalrymple Trough, in the Early and the upper Late
Miocene. In this framework, two regional angular unconformities, “U” and “M”, mark the episodes of
uplift and subsidence. Understanding the signification of these unconformities is critical to the
understanding of the formation of the Dalrymple Trough and Murray ridge, and how it relates with the
history of the OFZ.
2.3. Regional unconformities in the Arabian Sea
U-unconformity and the uplift of the Murray Ridge
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
5
The U-unconformity is recognized over the entire Owen Basin and the Oman abyssal plain (Shipboard
Scientific Party, 1974; Whitmarsh et al., 1974; 1979; Rodriguez et al., 2014). According to ODP
(Ocean Drilling Project) and DSDP (Deep Sea Drilling Project) Sites, the U-unconformity marks a
highly diachronous transition between Upper Oligocene-Lower Miocene pelagic chalk and Upper
Oligocene-Lower Miocene turbidites coming from the Indus fan. The U-unconformity does not have a
tectonic origin, and simply reflects the transition from Oligocene pelagites to Lower Miocene
turbidites as the substratum of the Owen Ridge gets progressively buried under the Indus deep-sea fan
(Shipboard Scientific Party, 1989; Mountain and Prell, 1990). Therefore, the angular unconformity
marking the main uplift of the Murray Ridge recognized by Gaedicke et al. (2002) (picked in blue on
Fig. 3) is different from the U-unconformity recognized everywhere in the Owen Basin and the Oman
abyssal plain (Rodriguez et al., 2014).
According to an industrial well located in the Indus abyssal plain (Pak-G2-1), the unconformity
marking the uplift of the Murray Ridge is ~8-10 Ma-old (Fig. 3) (Calvès, 2008). The latter
unconformity is formed by 8-9 Ma-old channel-levee system adopting an onlap configuration over a
pre-10 Ma-old tilted series of channel-levee systems (Kolla and Coumes, 1987; Calvès, 2008).
This age contrasts with earlier estimations based on a correlation with the age of uplift of the Owen
Ridge (Gaedicke et al., 2002), first assessed at 15-20 Ma (Mountain and Prell, 1990). Actually, the
East Oman Margin, the Owen Basin, and the Owen Ridge show compressive deformation precisely
dated at 8.2-8.8 Ma at ODP Site 730 (Rodriguez et al., 2014) consistent with the age of the
unconformity marking the uplift of the Southern Murray Ridge. An erosive surface at the top of the
Owen Ridge, marked by large submarine failures (Rodriguez et al., 2012; 2013), and dated at 8-9 Ma
by ties with ODP Site 722, characterizes the younger uplift of the Owen Ridge (Rodriguez et al.,
2014).
M-unconformity
The M-unconformity in the Oman abyssal plain marks both an episode of subsidence in the Dalrymple
Trough, and an abrupt tilt of an uniform, 4 km-thick sequence composed of Indus channel-levee
systems lying on the subducting plate (Fig. 4) (Gaedicke et al., 2002; Smith et al., 2012). The
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
6
overlying sequence adopts a wedge-shape geometry, with a maximal thickness of 3.5 km (Smith et al.,
2012). It is composed of turbiditic deposits coming from the Makran that onlap the M-unconformity
(Fig. 4). The latter is probably coeval with the last stage of uplift of the Murray Ridge, and the opening
of the Dalrymple Trough (Gaedicke et al., 2002).
This unconformity has never been drilled, which results in strong uncertainties in its age (Schlüter et
al., 2002). First related to a Messinian uplift event in the Zagros Mountain (Ross et al., 1986), it has
also been related to the onset of seafloor spreading in the Gulf of Aden (Schlüter et al., 2002),
estimated at 13 Ma when these studies were published (Cochran, 1981). The latter age cannot be
considered valid anymore according to recent magnetic anomalies studies that document the onset of
seafloor spreading in the Gulf of Aden as early as ~20 Ma (Fournier et al., 2010). Changes in the
regime of deformation in the Zagros suggest a 3-7 Ma-old kinematic change related to the Arabia-
Eurasia collision (Allen et al., 2004; Mouthereau et al., 2012) that may account for the M-
unconformity, but the existence of this kinematic change remains a matter of debate (Hatzfeld and
Molnar, 2010).
2.4. The Indus deep-sea fan
Tectonic deformation along the India-Arabia plate-boundary is well recorded by sediments belonging
to the Indus turbiditic system. At its thickest part the fan is more than 9-km thick, but its thickness
decreases when approaching the Owen-Murray Ridge (Fig. 3c) (Coumes and Kolla, 1984; Clift et al.,
2001). It forms a typical mud-rich, “passive margin fan” (sensu Reading and Richards, 1994), with
numerous inter-bedded pelagic layers (Shipboard Scientific Party, 1989). Indus Fan sedimentation
started during the Middle Eocene as the result of the onset of the India-Eurasia collision and
accelerated since the Early Miocene, coincident with a sharp increase in sedimentation rates related to
the uplift of the Himalaya and the onset of the Asian monsoon (Clift et al., 2001; Clift and Gaedicke,
2002; Clift et al., 2008). Seismic lines collected in the Indus deep-sea fan document the appearance of
well-developed channel-levee complexes since the Middle Miocene (Clift et al., 2001; Calvès, 2008).
In the Early Pleistocene, the Indus canyons underwent a southeast migration, leading to a major
episode of avulsion and concentration of Indus deposits on the southeastern part of the fan (fossil
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
7
Indus canyons are mapped in Figure 1) (Kolla and Coumes, 1987; Bourget et al., 2013). This
migration resulted in dominantly pelagic deposits along the OFZ, allowing a good preservation of the
fault scarps morphology on the seafloor (Shipboard Scientific Party, 1974; Rodriguez et al., 2011).
3. Material and Methods
The new dataset presented in this study was acquired onboard the French Navy oceanographic vessel
Beautemps-Beaupré during the OWEN1 and 2 (2009 and 2012) surveys. Multibeam bathymetry was
collected using a Kongsberg-Simrad EM 120 echosounder (Fig. 2), and combined with previously
published data acquired during the MARABIE (Bourget et al., 2010) and CHAMAK cruises (Ellouz-
Zimmerman et al., 2007a,b). Seismic reflection profiles of the OWEN2 cruise were acquired at 10
knots using two GI air-guns (one 105/105 c.i. and one 45/45 c.i., fired every 10 seconds at 160 bars in
harmonic mode, resulting in frequencies ranging from 15 to 120 Hz) and a 24-channel, 600 m-long
streamer. Seismic profiles have a common mid-point spacing of 6.25 m and achieved a sub-surface
penetration of ~2s TWT. The processing consisted in geometry setting, water-velocity normal move-
out, stacking, water-velocity F-k domain post-stack time migration, bandpass filtering and automatic
gain control. All profiles are displayed with a vertical exaggeration of 8 at the seafloor.
Two-way travel time to seismic reflectors was converted to depth using a P-wave velocity between
1530 and 1730 m. s
-1
for lower and upper bounds. This range of values covers safely the
measurements performed in the area in the same pelagic sediments (Shipboard Scientific Party, 1974,
1989). The highest P-wave velocity of 1950 m.s
-1
has been measured in the Pliocene turbidites
underneath the pelagic cover (White and Klitgord, 1976; Kolla and Coumes, 1987). The reflectors
picked on seismic profiles have been selected based on seismic discontinuities that reflect lithological
changes, stratigraphic hiatuses or tectonic deformation. In the following, sedimentary series before the
opening of the Dalrymple Trough are referred to as the “substratum” of the Dalrymple Trough, which
is employed in the sense of “pre-rift” series, in order to avoid the confusion with the continental
basement of the Murray Ridge observed on several profiles.
4. Stratigraphic framework
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
8
4.1. Criteria for the identification of sedimentary deposits
Turbiditic channels are characterized on seismic profiles by a typical lens-like architecture with a
concave-up lower boundary, and discontinuous, high amplitude reflection. The associated levees
display a wedge shape, with high amplitude, transparent seismic facies. Mass transport deposits
display the same chaotic to transparent seismic facies, but their geometry is more irregular. On the
other hand, pelagic deposits display well-stratified, continuous and conformable horizons on seismic
profiles. It is sometimes difficult to discriminate between turbiditic and pelagic deposits on seismic
profiles. Bottom-currents may influence the geometry of pelagic deposits, leading to typical pinched-
out, sigmoid geometries referred as contouritic drifts (Faugères et al., 1999). Figure 5 summarizes the
main criteria of sedimentary deposits identification.
4.2. Sedimentation rates
During turbiditic deposition (Late Miocene-Pliocene), the sedimentation rates ranged between 350 and
600 m Ma
-1
according to estimations at DSDP Site 222 located at the edge of the OFZ (Fig. 1; latitude
~20°N) (Shipboard Scientific Party, 1974). Turbiditic channels are sealed by a Plio-Pleistocene
pelagic drape according to ties with DSDP Site 222 (Shipboard scientific party, 1974; Rodriguez et al.,
2011; 2013). Several DSDP and ODP drillings are available along the Owen Ridge (Shipboard
Scientific Party, 1974; 1989), but the complete sedimentary sequence of the Murray Ridge has never
been drilled down to the basement (Schulz et al., 1998; Ziegler et al., 2010). All average Pleistocene
pelagic sedimentation rates estimated at different drilling sites along the Owen-Murray Ridge range
between 30 and 55 m Ma
-1
(Shipboard scientific party, 1974, 1989, and core KS07 in Bourget et al.,
2013), with the exception of rates of the order of 100 m Ma
-1
(Table 1) estimated from core MD 04873
(Ziegler et al., 2010). However, the nearby core S090-93KL (Schulz et al., 1998) (located only 40 km
away from the core MD 04873) (Fig. 2a) documents rates of about 50 m Ma
-1
, which suggests that
core MD 04873 has undergone core overpull (Skinner and McCave, 2003). There is consequently little
spatial variation of the Pleistocene pelagic sedimentation rates along the Owen Ridge, allowing large-
scale interpolation of these values in areas exclusively covered by pelagic deposits. However, these
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
9
values cannot be extrapolated in areas where bottom-currents seem to have interacted with pelagic
processes, as observed at the edge of the OFZ and the Qalhat Seamount (Fig. 8, 9).
The thickness of the pelagic sediments overlying a reflector marking a geological event can be
converted into time using uniform sedimentation rates, therefore providing an age estimate of the
geological event. Two major sources of uncertainties are inherent to this approach. First, the estimated
pelagic thickness depends on the value of P wave velocity used to convert two way travel time
distance into meters. We estimate different pelagic thicknesses using P wave velocities ranging
between 1550 and 1730 m.s
-1
(table 1). The second source of uncertainty is the regional variability of
sedimentation rates, ranging between 30m.Ma
-1
and 55 m.Ma
-1
(Table 1). The conversion of the
pelagic thickness into time is done for the upper and lower estimates of the sedimentation rates,
providing the widest range of acceptable age estimates (Table 1). A last source of uncertainties is the
measurement of pelagic thickness on seismic profiles, considered in the order of 10 ms (TWT). Taken
all together, the uncertainties related to each age mentioned hereafter are of the order of 1 Myr for the
Plio-Pleistocene interval, providing valuable constrains at the time scale of tectonic processes.
5. Structure of the Dalrymple horsetail
The Southern part of the Dalrymple Trough is an asymmetric structure comparable to a half-graben
oriented N50°E bordered by a single normal fault to the southeast, and a complex set of antithetic
faults to the west (Fig. 1) (Edwards et al., 2000). The connection between the OFZ and the Dalrymple
Trough forms a complex horsetail structure (Fig. 2) composed of several normal faults trending
perpendicular to the OFZ, hereafter referred as "transverse faults" (sensu Ben Avraham and ten Brink,
1989). Focal mechanisms indicate a minor strike-slip component in transverse faults motion (Fig. 2).
Transverse faults are in the continuation of a dense network of right-stepping, en-échelon faults on the
northwestern side of the trough (labelled 1 on Fig. 2). Transverse faults delineate subsiding sub-basins
within the trough, whose lengths range between 10 and 20 km (Fig. 2). The trough forms a syncline
basin on seismic lines (Fig. 3, 6) (Edwards et al., 2000; Gaedicke et al., 2002 a,b). Although variable
in complexity, this syncline pattern is identified from sub-basin to sub-basin. In detail, the sedimentary
sequences that form the syncline basin display a series of angular unconformities (Fig. 6, 7), but the
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
10
syncline itself is isopach, i.e. shows the same thickness throughout. Locally, a dense network of
transverse faults offset the syncline basin (Fig. 8). In the area of connection with the trough, the OFZ
is oriented N20°E, and forms a positive flower structure on seismic profiles, expressed by a pressure
ridge on the seafloor (Fig. 9, 10). There, contouritic drifts (Fig. 8, 10) mark the transition from
turbiditic to pelagic processes possibly related to a major avulsion episode of the Indus deep-sea fan
(Kolla and Coumes, 1987; Bourget et al., 2013). Since this transition, contouritic bodies at the edge of
the OFZ indicate the presence of a topography driving the bottom-current course.
At the entrance of the trough, the OFZ is bounded to the east by a set of ~N40°E en-échelon faults
(labelled 2 on Fig. 2), which merges northwards with the single normal fault bounding the eastern side
of the trough. Both the OFZ and the en-échelon fault system (2) cross cut a buried system of folds
observed on the seismic profiles (Fig. 9, 10). Two anticlines separated by a syncline are observed on
the eastern side of the OFZ (Fig. 9), whereas a syncline structure is observed on its western side (Fig.
10). The amplitude of the fold-system ranges between 1s (TWT) (Fig. 10) and 2s (TWT) (Fig. 9),
indicating bathymetric highs of the order of 700-1500 m before turbiditic covering. The top of one of
these anticlines is still forming a 700-m high arcuate relief at the latitude of 21°35'N (Fig. 2; 6). The
tilt of sedimentary layers on the eastern side of the arcuate relief indicates that some compression is
still active in this area (Fig. 9) and may represent an analog of the successive unconformities observed
in the sequence that was subsequently folded in the Dalrymple Trough (Fig.6). The system of folds
shows an isopach deformation at depth, followed by a gentle fanning configuration upward indicating
syn-tectonic deposition (Fig. 10). The seismic profile displayed in Fig. 10 highlights a contrast in the
deformation pattern on both sides of the OFZ, the syncline structure being observed solely on its
western side. The latter results from the right-lateral offset of the fold system, which moved the fold
initially located around the latitude of the profile in Fig. 10 to the latitude of the profile in Fig. 9.
The southern Murray Ridge follows a N60°E trend, and displays an asymmetric shape on seismic
sections, with a steeper flank facing the Dalrymple Trough (Fig. 3) (Edwards et al., 2000). The angular
unconformity marking the uplift of the southern Murray Ridge (picked in blue) is well identified on its
eastern side (Fig. 3) (Edwards et al., 2000; Clift et al., 2001; Gaedicke et al., 2002 a,b) and on its
western side (Fig.8).
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
11
6. Age estimates of the deformation
Opening of the Dalrymple Trough, Northern Murray Ridge uplift, and M-unconformity
The age of opening of the Dalrymple Trough and of related geological events is estimated by using
turbiditic channel-levee systems as morphological markers of the deformation and by regional
correlations with ODP Site 722 at the Owen Ridge (Fig. 1).
Previously published multibeam bathymetric data reveal the presence of fossil meandering turbiditic
channels on the top of the Murray Ridge (Fig. 2) and the Jinnah High (Ellouz-Zimmermann et al.,
2007, Mouchot, 2009; Mouchot et al., 2008, 2010). Turbiditic channels are identified below ~4.1 s
(TWT) in the vicinity of the OFZ (Fig. 9, 10) and on the western flank of the Dalrymple Trough (Fig.
6). Perched turbiditic channels are also observed on the eastern flank of the trough (Fig. 6, 8).
All turbiditic channels in the area are blanketed by pelagites that preserved their seafloor expression,
but none of them are active anymore in the vicinity of the trough (Deptuck et al., 2003; Ellouz-
Zimmermann et al., 2007a,b; Rodriguez et al., 2011). Because of the steep and uneven slopes created
by the active thrusts of the Makran wedge, the turbiditic canyons cutting through the wedge do not
form any turbiditic channel at their mouth (Bourget et al., 2011). North of the Dalrymple Trough, the
Jinnah High acts as a topographic barrier for the turbiditic deposits coming from the Makran (Fig. 4).
The origin of the turbiditic sequence observed on both sides of the Dalrymple Trough and below the
M-unconformity is thus the Indus deep-sea fan (Gaedicke et al., 2002a,b; Ellouz-Zimmermann et al.,
2007a,b). Therefore, the youngest generation of channel-levee system in the area pre-dates the opening
of the Dalrymple Trough and the formation of the M-unconformity, and gives their maximal age. The
age of abandonment of the youngest channel-levee is estimated by the thickness of the overlying
pelagic cover, converted into time using uniform pelagic sedimentation rates in the area (30-55
m Ma
-1
).
Channel-levee systems (labelled A) observed both on the top of the Jinnah High (Ellouz-Zimmerman
et al., 2007) and west of the Dalrymple Trough (Fig. 6) are covered by ~0.3 s (TWT) of pelagic
deposits. Channel-levee systems A are therefore ~5.8 ± 2.2 Ma old (Table 1b). On the other hand, the
channel-levee system (labelled B on Figures 6 and 8) observed on the west side of the connection
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
12
between the OFZ and the Dalrymple Trough is covered by a ~0.18 s TWT-thick pelagic cover, making
this channel the youngest in the area. This pelagic thickness indicates that turbiditic sedimentation
stopped in the area since at least 3.5 ± 1.4 Ma (Table 1b). This channel-levee system is dissected by
the transverse faults at the entrance of the trough, indicating that it predates this episode of
deformation (Fig. 8). The maximal plausible age of the Dalrymple Trough is therefore 3.5 ± 1.4 Ma.
Channel-levee systems in the area migrated southwards prior to their deactivation 3.5 ± 1.4 Ma ago,
which indicates that the Oman abyssal plain seafloor was not tilted towards the subduction zone before
that time (Mouchot, 2009).
The age of opening of the Dalrymple Trough can be refined by seismic correlations performed on the
profile transverse to the trough (Fig. 6, 7). The syncline basin forming the trough displays a seismic
facies typical of a turbiditic and pelagic layers succession (Fig. 6, 7). It further indicates that the
turbiditic sequence identified on both sides of the trough used to be connected before its opening (Fig.
6-8). An unconformable ponded sedimentary sequence, mainly composed of mass transport deposits,
seals the syncline basin (Fig. 7). The turbiditic series deformed by the syncline thus corresponds to the
substratum of the Dalrymple Trough, the top of which corresponds to the last sedimentary layer
deposited before its opening. The uppermost pelagic reflector of the syncline (Fig. 6, 7) correlates
fairly well with the sedimentary sequence west of the trough, where it is overlain by a 0.1 s (TWT)-
thick pelagic cover corresponding to 1.9 ± 0.9 Myr. The most probable age of opening of the
Dalrymple Trough is therefore 1.9 ± 0.9 Ma. About 1200-m of subsidence has been accommodated by
the trough since 1.9 ± 0.9 Ma.
On the other hand, the M-unconformity observed on Figure 4 can be correlated up to the Jinnah High.
Although indicating a tectonic tilt of the Indus channel-levee systems in the Oman abyssal plain, the
dip of the unconformity slightly changes at the approach of the Jinnah High. This change in dip is due
to the progressive transition from pelagic to turbiditic deposits as the Makran turbiditic system
progressively buries the M-unconformity. The thickness of pelagic deposits covering the M-
unconformity in the area uncovered by Makran turbidites is 0.1 s (TWT), which gives an age of about
~1.9 ± 0.9 Ma. Since the reflector marking the opening of the Dalrymple Trough (Fig. 6) and the M-
unconformity are sealed by the same thickness of pelagic sediments, both events are considered
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
13
synchronous. Kolla and Coumes (1987) previously noticed the southeastward shift of the Pleistocene
Indus channel-levee systems that may relate to the Pleistocene uplift of the northern Murray Ridge..
Regional correlations of seismic lines show that the pelagic layer overlying the M-unconformity in the
vicinity of the Dalrymple Trough is also recognized in the vicinity of the 20°N pull-apart basin and at
the Southern Owen Ridge (Fig. 11). At the 20°N basin, the M-unconformity corresponds to a 2.5 ±1
Ma-old reflector marking the onset of contourite deposition on the western side of the basin
(Rodriguez et al., 2013b). There, the onset of contourite deposition may reflect local disturbance of
bottom currents following the opening of the basin (Rodriguez et al., 2013b). At the Owen Ridge, the
reflector corresponding to the M-unconformity is dated at 2.4 Ma at ODP Site 722 (Discoaster
pentaradiatus, sampled at a depth of 74 mbsf, Shipboard Scientific Party, 1989). The age of 2.4 Ma is
in the range of ages predicted by the extrapolation of sedimentation rates, and may represent the most
likely age of the last structural reorganization of the OFZ expressed by the coeval opening of both the
Dalrymple Trough and the 20°N Basin and the M-unconformity. Considering a steady India-Arabia
motion of 3±1 mm.yr
-1
(Fournier et al. 2008a,b; 2011), the Dalrymple Trough opened in response to a
limited amount of strike-slip motion, in the order of 5-10 km.
Late Miocene Murray Ridge uplift and buried folds
The use of the distribution of turbiditic channels as a marker of the evolution of the Murray Ridge
elevation is more problematic. The profile displayed in Fig. 3c shows that before its 1.9± 0.9 Ma
uplift, the northern Murray Ridge was almost totally buried by Indus turbidites. The paleo-location of
the Indus canyon in the northwestern extremity of the Indian margin prior to the Pleistocene (Fig.1,
Kolla and Coumes, 1987) implies that turbiditic channels simply bypassed the southern Murray Ridge
to flood the Oman abyssal plain. The southern Murray Ridge was thus a prominent high since 8-10 Ma
(age of the unconformity marking its uplift; Calvès, 2008), but did not act as a major barrier for the
Indus turbidites due to the paleo-location of the canyon (Fig. 1). The northern Murray Ridge, although
uplifted around 8-10 Ma (Fig. 4), was less prominent than the Southern Murray Ridge and had been
rapidly buried under Indus turbidites. The age of the buried folds identified on Figures 9 and 10 cannot
be estimated because of the large range of Late Miocene-Pliocene Indus sedimentation rates (Kolla
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
14
and Coumes, 1987; Clift et al., 2001; Calvès, 2008), but they are probably coeval with the main uplift
episode of the southern Murray Ridge at 8-10 Ma. The N-S trend of the arcuate relief at ~21°35°N
implies a roughly E-W compression at the origin of the Late Miocene deformation.
7. Discussion
7.1. Mode of opening of the south Dalrymple Trough
The main structural characteristic of the Dalrymple Trough is the numerous transverse faults that form
the horsetail structure. Transverse faults are common structures within narrow (<15-km-wide)
stepover areas such as the Dead Sea Basin (Kashai and Croker, 1987; Ben Avraham and ten Brink,
1989; Lazar et al., 2006; Smit et al., 2008) or the 20°N Basin along the OFZ (Rodriguez et al., 2011,
2013). There, transverse faults form during the opening of the pull-apart basin and were shown to
transfer strike slip motion from one main bounding strike-slip segment to the other (Ben Avraham and
ten Brink, 1989). Stepovers at the origin of the Dalrymple Trough (>30 km), the North Aegean Sea
(>40 km), the Andaman Sea (>100 km) are far much wider, and emplaced in areas of strong crustal
thickness variations inherited from previous geological events. Complex structures such as the Jinnah
Seamount (Dalrymple), the North Cycladic Detachment Fault (Aegean Sea; Le Pourhiet et al., 2012;
Jolivet et al., 2013) or the Alcock Rise (Andaman Sea; Morley, 2013) prevent any connection between
bounding strike-slip faults at the extremities of the stepover. In this context, transverse faults cannot
result from the transfer of strike-slip motion between bounding strike-slip segments. Analog modeling
works (Basile and Brun, 1999) propose that in wide stepover areas, a divergent system composed of
normal faults trending perpendicular to the main strike-slip direction emplaces in the first stages of
strike-slip motion. As the amount of strike-slip motion increases, the main strike-slip fault connects
progressively the transverse normal faults. The first step of these analog models strikingly reproduces
the en-échelon fault system (2) observed at the Dalrymple Trough (Fig. 12). However, the full
connection of the strike-slip fault with the transverse normal faults implies a large amount of relative
motion in the model, which contrasts with the case of the Dalrymple Trough, where only ~5-10 km of
relative motion were accommodated since the first stages of opening ~1.9 ± 0.9 Ma. The ~5-10 km of
relative motion at the Dalrymple Trough drastically contrasts with its dimensions (150-km-long, 30-
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
15
km-wide). This raises the question of the relationship between the size of a basin and the amount of
finite motion. The formation of a stepover in response to the structural reorganization of the strike-slip
fault system isolated a subsiding half graben of dimensions close to the present-day Dalrymple
Trough. The connection between the main strike-slip fault and the numerous transverse normal faults
may occur in the earliest stages of basin opening, and not progressively as suggested by analog
models. The numerous transverse faults individualized several sub-basins, and subsequently
accommodated distributed transtension corresponding to the ~5-10 km of India-Arabia relative
motion.
Moreover, the opening of the Dalrymple Trough enhanced the uplift of its flanks, and rejuvenated the
topography at the northern Murray Ridge and the westernmost part of the southern Murray Ridge,
similar to what is observed along the Dead Sea flanks (Basile and Allemand, 2002). On the other
hand, the southern Murray Ridge is outside the area affected by the flexure induced by the Dalrymple
Trough, and has probably been little affected by this uplift phase (Fig. 1d).
7.2. Tectonic history of the Dalrymple Trough-Murray Ridge system
The results above document at least two major deformation events (Fig. 13) : a first compressional one
(8-10 Ma-old) related to the folds observed on the sides of the OFZ (Fig. 9, 10) the main uplift of the
southern Murray Ridge, and the first uplift of the Northern Murray Ridge (Fig. 3) (Kolla and Coumes,
1988; Calvès, 2008); and a second one (1.9 ± 0.9 Ma-old) related to the subsidence of the Dalrymple
Trough, the uplift of its flanks (including the major uplift of the northern Murray Ridge), and the
formation of the M-unconformity in the Oman abyssal plain (Fig. 4, 6, 7). Compressive deformation
was still active in the area since 8-10 Ma. Whether stepover basins inception is coeval with the
inception of the OFZ is difficult to assess, as some stepover basins were shown to develop a few
million years after the initiation of the San Andreas Fault (e.g. Wakabayashi, 2007) and the Levant
Fault (Garfunkel and Ben Avraham, 2001). The alternative involves a third, intermediate episode, in
which the OFZ emplaced at ~6 Ma, and then underwent opening of stepover basins at 1.9 ± 0.9 Ma.
However, the location of the plate boundary for the 3-6 Ma time span is unknown. It is even possible
than India and Arabia were temporary coupled during this time span, both Carlsberg and Sheba ridges
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
16
showing similar half spreading rates at ~3 cm.yr
-1
(Merkouriev and DeMets, 2006; Fournier et al.,
2010). This framework, summarized in Figure 13, contrasts with the two stages history (in the Early
and the Late Miocene) previously assessed (Gaedicke et al., 2002), and implies a new geodynamic
interpretative scheme of the Dalrymple Trough-Murray Ridge system.
Origin of the ~8-10 Ma-old compressional episode of deformation
The uplift of the Southern Murray Ridge is coeval with widespread deformation in the Owen Basin
(Rodriguez et al., 2014) and the ending of seafloor spreading rates deceleration at 8-11 Ma recorded
by magnetic anomalies at the Carlsberg Ridge (Mercuriev and DeMets, 2006). The Late Miocene
deformation in the Arabian sea coincides in time with a Late Miocene global plate reorganization
event, which is expressed in the Indian Ocean by the separation of Somalia from Africa (DeMets et al.,
2005; Mercuriev and DeMets, 2006) and intra-plate deformation in the Central Indian Ocean,
separating India from Australia (Weissel et al., 1980; Wiens et al., 1985; Bull and Scrutton, 1990;
1992; Chamot-Rooke et al., 1993; Henstock and Minshull, 2004; Delescluse et al., 2008; Krishna et
al., 2009; Copley et al., 2010). The increase in stress applied by the Miocene growth of the Himalayas
on India's plate boundaries has been tentatively proposed as the driver for this plate reorganization
event (Molnar et al., 1993; Molnar and Stock, 2009; Bull et al., 2010). However, the present-day
deformation west of the Chagos-Laccadive is extension rather than compression (Henstock and
Minshull, 2004). The compressive deformation at the India-Arabia plate boundary probably results
from complex tectonic interactions between both plates.
Origin of the ~1.9 ± 0.9 Ma-old transtensive episode of deformation
The M-unconformity records both the tilt of Indus series in front of the Makran and the opening of the
Dalrymple Trough, suggesting a common origin. This episode of deformation induced a second uplift
of the northern Murray Ridge. The opening of the Dalrymple Trough isolated the Makran from Indus
sediments, the only remaining source of sediments being the Makran turbiditic system. The younger
age of the M-unconformity implies that the erosion rates of the Makran accretionary wedge have been
strong during the Pleistocene, probably in the same order as current rates (>2 mm.yr
-1
, Bourget et al.,
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
17
2011). Sedimentation rates control the temperature at the deformation front, high sedimentation rates
favoring shallow seismogenic rupture along megathrusts (Smith et al., 2012; 2013). This Late Pliocene
transition in the sedimentation regime may have affected both the structural organization of the wedge
and its seismogenic potential.
The reflector marking the opening of the Dalrymple Trough and the 20°N Basin being the same (Fig.
11) , the reorganization of the OFZ between 20°N and 24°N may be coeval with the last structural
reorganization of the Makran accretionary wedge marked by the M-unconformity.
8. Conclusions and perspectives
This study documents a Late Miocene compressive episode along the India-Arabia plate boundary
expressed by a complex set of folds at the connection between the OFZ and the Dalrymple Trough.
The U-unconformity is not recognized at the Murray Ridge, whose uplift is recorded by an angular
unconformity dated at 8-10 Ma (Calvès, 2008). A Late Miocene compressive episode of deformation
is recognized throughout the Owen Basin (Rodriguez et al., 2014) and is linked to a Late Miocene
global plate reorganization event.
This study also shows the formation of a stepover area ~1.9 ± 0.9 Ma at the origin of both the
Dalrymple Trough and the second uplift of the Northern Murray Ridge. This age is significantly
younger than the Miocene age previously proposed (Gaedicke et al., 2002) and in better agreement
with the recent age of the OFZ, which shows no structure older than 3 Ma along strike (Rodriguez et
al., 2011; 2013). This younger age also implies that large stepover basins (150-km-long) can develop
with limited amount of strike-slip motion (5-10 km). Moreover, the M-unconformity is not coeval with
the structural reorganization that affected the Makran wedge during Tortonian (McCall, 1997; Burg et
al., 2008; Smit et al., 2010), and marks a younger episode of structural reorganization 1.9 ± 0.9 Ma.
The reassessment of the tectonic framework of the Dalrymple Trough and the Murray Ridge points out
interesting regional perspectives for the Indus turbiditic system and the Makran accretionary wedge.
The last structural reorganization marked by the M-unconformity is roughly coeval with the opening
of the Dead Sea Basin along the Levant Fault (TenBrink et al., 1989). Whether coeval deformation
events along both strike-slip boundaries of the Arabian plate reflect a poorly constrained Late
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
18
Pliocene-Early Pleistocene kinematic change remains enigmatic (Allen et al., 2004; Smit et al., 2009,
Schattner, 2010). Considering the regional correlation with ODP Site 722, the last structural
reorganization marked by the M-unconformity is likely to be synchronous with a major monsoon
intensification over Asia around 2.4 Ma (Bloemendal and DeMenocal, 1989; An et al., 2001; Wang et
al., 2005; Huang et al., 2007), affecting the Makran and the Himalayan foreland. The critical Coulomb
wedge theory and geological studies (Berger et al., 2008; Whipple, 2009; Malavieille, 2010;
Iaffaldano et al., 2011) suggest that major climate changes can affect the wedge taper at a regional
scale, and induce widespread structural reorganization. Deep-sea drillings would be useful to decipher
the precise timing of the deformation in the area and confirms the link with monsoon intensity.
Second, the uplift of the northern Murray Ridge since ~1.9 ± 0.9 Ma might account for the Pleistocene
southwestward migration of the Indus canyon on the Indian Margin (Kolla and Coumes, 1987). A
precise dating of the episodes of migration of the canyon, together with reconstructions of the
evolution of the topography fossilized by channel-levee systems, would highlight the complex
interaction between tectonic and sedimentary processes (see similar examples in Mulder et al., 2012).
Acknowledgements
We thank the SHOM and GENAVIR team for their help in data acquisition. Processing of the OWEN-
2 dataset was carried out using the Geocluster 5000 software from CGGVeritas. We thank P. Dubernet
and N. Bacha for technical assistance, and L. Le Pourhiet for scientific discussions about horsetail
terminations. Tim Minshull and Lisa McNeill provided very detailed and constructive comments that
greatly helped us to improve this article. This study was supported by SHOM, IFREMER, INSU-
CNRS, and CEA (LRC Yves-Rocard).
Figure captions :
Figure 1 :a) Shaded-relief bathymetry of the Arabian sea and surrounding topography, from Becker et
al., 2009 and Fournier et al., 2011. LMR : Little Murray Ridge. Red stars: drill sites b) Simplified
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
19
cross section of the Oman abyssal plain, modified from White and Klitgord (1976) (see a) for
location). c) and d) : topographic profiles running transverse to the northern and southern Murray
Ridge. Inset shows the simplified geodynamic framework of the Arabian Sea. AOC: Aden-Owen-
Carlsberg triple junction.
Figure 2 : Multibeam bathymetric map of the Dalrymple Trough (a), and interpretative morpho-
structural scheme (b), with local crustal seismicity since 1973 (focal depth < 50 km, magnitude
Mw > 2) from available databases. The small circle about the closure-enforced MORVEL Arabia–
India rotation pole (dark blue dashed line; DeMets et al., 2010) is parallel to the trend of the OFZ at
the entrance of the trough, whereas the small circle determined from the active trace of the OFZ (light
blue dashed line, Fournier et al., 2011) is parallel to the trace of the OFZ up to 21°30°N and to the en-
échelon fault system labeled “2”. Location of seismic profiles is indicated by black dashed lines. Insets
show close bathymetric views of the turbiditic channels observed in the area, labeled A and B
according to their age of activity (A : older channel, 5.8 ± 2.2 Ma, B : younger channel, 3.5 ± 1.4 Ma).
Yellow stars: drill sites.
Figure 3 : a-b) Line drawing of previously published seismic profiles and c) seismic line crossing in
the Dalrymple Trough-Murray Ridge area(see the inset in the upper right hand corner for location and
Fig. 2). a) Line drawing of a profile crossing the southern Murray Ridge, from Gaedicke et al., 2002b.
The U-unconformity is not recognized on the eastern side of the Dalrymple Trough and in the vicinity
of the Murray Ridge; b) Line drawing of a profile crossing the Dalrymple Trough and the southern
Murray Ridge, modified after Edwards et al., 2000. A potential fracture zone offset is identified on the
eastern flank of the Murray Ridge; c) Profile modified from Kolla and Coumes (1987) showing an
angular unconformity on the eastern side of the Murray Ridge. The angular unconformity was dated at
8-10 Ma by calibration with an industrial drilling located in the Indus deep-sea fan (Pak-G2-1; Calvès,
2008). CLS= Channel-Levee Systems.
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
20
Figure 4 : Seismic profile from the Chamak cruise (Ellouz-Zimmerman et al., 2007; Mouchot, 2009)
crossing the Jinnah High, showing the M-unconformity (see figure 2 for location).a) close view of the
M-unconformity in the area of the Jinnah High; b) Seismic profile showing the M-unconformity and c)
its line drawing. CLS = channel-levee system.
Figure 5 : Synthesis of the seismic facies of the main sedimentary deposits observed in the study area.
Figure 6 : a) Bathymetry of the connection between the Owen Fracture Zone (OFZ) and the
Dalrymple Trough, showing en-échelon faults connecting transverse normal faults within the trough.
b) Seismic profile transverse to the Dalrymple Trough (see Fig. 2 and 6a for location) and c)
interpretation. MTD = mass transport deposits. CLS = channel-levee system.
Figure 7 : Close view of the seismic profile displayed in fig. 6, showing the stratigraphic correlation
of the last deformed layer within the trough with its western border. The last deformed layer within the
trough coincides with the M unconformity in the Oman Abyssal Plain. Close views of the profile
highlight the similar seismic sequence within and outside the Dalrymple Trough prior its opening.
Figure 8 : a) seismic profile crossing the transverse fault system and b) the interpretation.
Figure 9 : a) seismic profile crossing the Owen fracture Zone at the entrance of the Dalrymple Trough
(see Fig. 2 and 6 for location) and b) the interpretation.
Figure 10 : a)Seismic profile crossing the Owen Fracture Zone (OFZ) and the en-échelon fault system
2 (see Fig. 2 and 6 for location) and b) its interpretation. c) shows a bathymetric view of the en-
échelon fault system 2, with a particular emphasis over its relationship with fossil Indus turbiditic
channels.
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
21
Figure 11 : Regional correlation of seismic facies at the edges of the Dalrymple Trough, the 20°N
Basin (Rodriguez et al., 2013) and the Owen Ridge (Shipboard Scientific Party, 1989; Rodriguez et
al., 2014).
Figure 12 : Simplified structural sketchmap of the Dalrymple Trough and comparison with results
from analog modeling experiments from Basile and Brun (1999).
Figure 13 : Sketches of the geological history of the Dalrymple Trough and the Murray Ridge. OFZ :
Owen Fracture Zone, PB : Plate-Boundary. Stage A : A Late Miocene episode (8-10 Ma) of
compression along the India-Arabia plate boundary triggered the main uplift of the Murray Ridge. The
plate boundary was probably located on the eastern side of the Murray Ridge according to sharp
offsets of the basement observed on previously collected seismic lines (see Fig. 3). Stage B : This
intermediate stage considers the alternative in which the opening of the Dalrymple Trough is not
coeval with the inception of the OFZ. Since the emplacement of the OFZ between 3-6 Ma, Late
Miocene folds were offset dextrally and progressively buried under Indus turbidites. Stage C: The
Dalrymple Trough opened (~2-3 Ma), and provoked a new episode of uplift of the northern Murray
Ridge. It is related to the formation of the M-unconformity in the Oman abyssal plain. At the same
time, the Indus canyon began its migration towards the south-east.
Table 1 : a) Range of sedimentation rates calculated at different drilling sites in the Arabian Sea (see
Fig. 1 for location). b) Age estimations of the different markers used in this study, from conversion of
pelagic sediment thicknesses into time. Different values of sedimentation rates calculated at the
different drilling/coring sites available, as well as different P-wave velocities, are used for the
conversion of the pelagic thickness into time.
References
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
22
Allen, M., Jackson, J., Walker, R., 2004. Late Cenozoic reorganization of the Arabia-Eurasia collision
and the comparison of short term and long term deformation rates. Tectonics 23, TC2008,
doi:10.1029/2003TC001530.
An, Z., Kutzbach, J.E., Prell, W.L., Porter, S.C., 2001.Evolution of Asian monsoons and phased uplift
of the Himalaya–Tibetan plateau since Late Miocene times. Nature 411, 62–66.
Basile, C., Brun, J.P., 1999. Transtensional faulting patterns ranging from pull-apart basins to
transform continental margins: an experimental investigation. Journal of Structural Geology 21,
23–37.
Basile, C., Allemand, P., 2002. Erosion and flexural uplift along transform faults. Geophysical Journal
International 151, 646-653.
Becker, J.J., Sandwell, D.T., Smith, W.H.F., Braud, J., Blinder, B., Depner, J., Fabre, D., Factor, J.,
Ingalls, S., Kim, S. H., Ladner, R., Marks, K., Nelson, S., Pharaoh, A., Trimmer, R.,
VonRosenberg, J., Wallace, G., Weatherall, P., 2009. Global bathymetry and elevation data at 30
arc second resolution: SRTM30PLUS. Mar. Geod. 32, 355–371.
http://dx.doi.org/10.1080/01490410903297766.
Ben Avraham, Z., Ten Brink, U. S., 1989. Transverse faults and segmentation of basins within the
Dead Sea Rift. Journal of African Earth Science 8, 603-616.
Berger, A., Gulick, S.P.S., Spolita, J. A., Upton, P., Jaeger, J. M., Chapman, J. B., Worthington, L. A.,
Pavlis, T. L., Ridgway, K. D., Willems, B. A., McAleer, R. J., 2008. Quaternary tectonic response
to intensified glacial erosion in an orogenic wedge. Nature geoscience 1, 793-799,
doi:10.1038/ngeo334
Bourget, J., Zaragosi, S., Ellouz-Zimmermann, N., Ducassou, E., Prins, M.A., Garlan, T., Lanfumey,
V., Rouillard, P., Schneider, J-L., 2010. Highstand vs. lowstand turbidite system growth in the
Makran active margin: imprints of high-frequency external controls on sediment delivery
mechanisms to deep water systems. Marine Geology 274, 187-208
Bourget, J., Zaragosi, S., Ellouz-Zimermann, N., Mouchot, N., Garlan, T., Schneider, J.-L., Lanfumey,
V., Lallemant, S., 2011.Turbidite system architecture and sedimentary processes along
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
23
topographically complex slopes: The Makran convergent margin. Sedimentology 58, 376–406,
doi:10.1111/j.1365-3091.2010.01168.x
Bourget, J., Zaragosi, S., Rodriguez, M., Fournier, M., Garlan, T., Chamot-Rooke, N., 2013. Late
Quaternary megaturbidites of the Indus Fan: Origin and stratigraphic significance. Marine Geology
336, 10-23, http://dx.doi.org/10.1016/j.margeo.2012.11.011.
Bull, J. M., Scrutton, R. A., 1990. Fault reactivation in the central Indian Ocean and the rheology of
oceanic lithosphere. Nature 344, 855-858.
Bull, J. M., Scrutton, R. A., 1992. Seismic reflection images of intraplate deformation, central Indian
Ocean, and their tectonic significance. J. Geol. Soc. London 149, 955-966.
Bull, J. M., DeMets, C., Krishna, K. S., Sanderson, D. J., Merkouriev, S., 2010. Reconciling plate
kinematic and seismic estimates of lithospheric convergence in the central Indian Ocean. Geology
38, 307-310, doi: 10.1130/G30521.1
Burg, J.P., Bernoulli, D., Smit, J., Dolati, A., Bahroudi, A., 2008. A giant catastrophic mud-and-debris
flow in the Miocene Makran. Terra Nova 20, 188–193, doi: 10.1111/j.1365-3121.2008.00804.x
Burgath, K. P., Von Rad, U., Van der Linden, W., Block, M., Ali Khan, A., Roeser, H. A., Weiss, W.,
2002. Basalt and peridotites from Murray Ridge: are they of supra-subduction origin? in The
Tectonic and Climatic Evolution of the Arabian Sea Region, edited by P. D. Clift et al., Geological
Society Special Publication, 195, 117-135
Cattin, R., Chamot-Rooke, N., Pubellier, M., Rabaute, A., Delescluse, M., Vigny, C., Fleitout, L.,
Dubernet, P., 2009. Stress change and effective friction coefficient along the Sumatra-Andaman-
Sagaing fault system after the 26 December 2004 (Mw=9.2) and the 28 March 2005 (Mw=8.7)
earthquakes. Geochemistry, Geophysics, Geosystems 10, doi:10.1029/2008GC002167.
Calvès, G., 2008. Tectonostratigraphic and climatic record of the NE Arabian Sea, Ph.D. thesis, 305
pp., Univ. of Aberdeen, Aberdeen U. K.
Chamot-Rooke, N., Jestin, F., DeVoogd, B., 1993. Intraplate shortening in the central Indian-ocean
determined from a 2100-km-long north-south deep seismic-reflection profile. Geology 21, 1043–
1046.
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
24
Chamot-Rooke, N., Fournier, M., Scientific Team of AOC and OWEN cruises, 2009. Tracking
Arabia-India motion from Miocene to Present. American Geophysical Union, Fall Meeting 2009.
Clift, P. D., Shimizu, N., Layne, G. D., Blusztain, J. S., Gaedicke, C., Schluter, H.-U., Clark, M. K.,
Amjad, S., 2001. Development of the Indus Fan and its significance for the erosionalhistory of the
Western Himalaya and Karakoram. Geological Society of America Bulletin 113(8), 1039–1051.
Clift, P. D., Gaedicke, C, 2002. Accelerated mass flux to the Arabian Sea during the middle to late
Miocene. Geology 30, 207-210. doi:10.1130/0091-13
Clift, P. D., Hodges, K. V., Heslop, D., Hannigan, R., Van Long, H., Calvès, G., 2008. Correlation of
Himalayan exhumation rates and Asian monsoon intensity. Nat. Geosci. 1, 875-880,
doi:10.1038/ngeo351
Cochran, J. R., 1981. The Gulf of Aden: structure and evolution of a young ocean basin and
continental margin. Journal of Geophysical Research 86, 263-287.
Copley, A., Avouac, J-P., Royer, J-Y., 2010. India-Asia collision and the Cenozoic slowdown of the
Indian plate: implications for the forces driving plate motions. J. Geophys. Res. 115, B03410,
doi:10.1029/2009JB006634
Coumes, F., Kolla, V., 1984. Indus Fan: seismic structure, channel migration and sediment thickness
in the upper fan. in: Haq, B.U., J.D., M. (Eds.), Marine Geology and Oceanography of Arabian Sea
and coastal Pakistan. Van Nostrand Reinhold Comp., New York, pp. 101-110.
Delescluse, M., Chamot-Rooke, N., 2007. Instantaneous deformation and kinematics of the India-
Australia Plate. Geophysical Journal International 168, 818-842, doi: 10.1111/j.1365-
246X.2006.03181.x
Delescluse, M., Montési, L. G. J, Chamot-Rooke, N., 2008. Fault reactivation and selective
abandonment in the oceanic lithosphere. Geophysical Research Letters 35, L16312
doi:10.1029/2008GL035066
DeMets, C., Gordon, R., Royer, J-Y., 2005. Motion between the Indian, Capricorn and Somalian
plates since 20 Ma: implications for the timing and magnitude of distributed lithospheric
deformation in the equatorial Indian Ocean. Geophysical Journal International 161, 445-468, doi:
10.1111/j.1365-246X.2005.02598.x
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
25
DeMets, C., Gordon, R.G., Argus, D.F., Stein, S., 1990. Current plate motions. Geophysical Journal
International 101, 425-478.
DeMets, C., Gordon, R.G., Argus, D.F., Stein S., 1994. Effect of recent revisions of the geomagnetic
reversal time scale on estimates of current plate motions. Geophysical Research Letters 21, 2191-
2194.
DeMets C., Gordon R. G., Argus, D.F., 2010. Geologically current plate motions, Geophysical Journal
International 181, 1-80. doi: 10.1111/j.1365-246X.2009.04491.x.
Deptuck, M. E., Steffens, G. S., Barton, M., Pirmez, C., 2003. Architecture and evolution of upper fan
channel-belts on the Niger Delata slope and in the Arabian Sea. Marine and Petroleum Geology
20, 649-676.
Edwards R.A., Minshull, T. A., White, R. S., 2000. Extension across the Indian–Arabian plate
boundary: the Murray Ridge. Geophysical Journal International 142, 461-477.
Edwards, R. A., Minshull, T. A., Flueh, E. R., Kopp, C., 2008. Dalrymple Trough: An active oblique-
slip ocean-continent boundary in the northwest Indian Ocean. Earth and Planetary Science Letters
272, 437-445.
Ellouz Zimmermann, N., Deville, E., Müller, C., Lallemant, S., Subhani, A. B., Tabreez, A. R., 2007a.
Impact of sedimentation on convergent margin tectonics : example of the Makran Accretionary
prism (Pakistan). Thrust Belts and Foreland Basins: From Fold Kinematics to Hydrocarbon
Systems, edited by O. L. Lacombe et al., pp. 327–350, Springer, Berlin.
Ellouz-Zimmermann, N. et al., 2007b. Offshore frontal part of the Makranaccretionary prism
(Pakistan) the Chamak Survey in Thrust Belts and Foreland Basins: From Fold Kinematics to
Hydrocarbon Systems, edited by O. L. Lacombe et al., pp. 349–364, Springer, Berlin.
Faugères, J.C., Stow, D.A.V., Imbert P., Viana, A. 1999. Seismic features diagnostic of contourite
drifts, Marine Geology 162, 1-38.
Ford, M., Rohais, S., Williams, E. A., Bourlange, S., Jousselin, D., Backert, N., Malartre, F., 2013.
Tectono-sedimentary evolution of the western Corinth rift (Central Greece). Basin Research 25, 3-
25, doi : 10.1111/j.1365-2117.2012.00550.x
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
26
Fournier, M., Patriat, P., Leroy, S., 2001. Reappraisal of the Arabia–India–Somalia triple junction
kinematics. Earth and Planetary Science Letters 189, 103–114.
Fournier, M., Chamot-Rooke, N., Petit, C., Fabbri, O., Huchon, P., Maillot, B., Lepvrier, C., 2008a.
In-situ evidence for dextral active motion at the Arabia–India plate boundary. Nature Geoscience 1,
54–58 http://dx.doi.org/10.1038/ngeo.2007.24.
Fournier, M., Petit, C., Chamot-Rooke, N., Fabbri, O., Huchon, P., Maillot, B., Lepvrier, C., 2008b.
Do ridge-ridge-fault triple junctions exist on Earth? Evidence from the Aden–Owen–Carlsberg
junction in the NW Indian Ocean. Basin Research 20, 575–590 http://dx.doi.org/10.1111/j.1365-
2117.2008.00356.x.
Fournier, M., Chamot-Rooke, N., Petit, C., Huchon, P., Al-Kathiri, A., Audin, L., Beslier, M.-O.,
d'Acremont, E., Fabbri, O., Fleury, J.-M., Khanbari, K., Lepvrier, C., Leroy, S., Maillot, B.,
Merkouriev, S., 2010. Arabia–Somalia plate kinematics, evolution of the Aden–Owen–Carlsberg
triple junction, and opening of the Gulf of Aden. Journal of Geophysical Research 115, B04102
http://dx.doi.org/10.1029/2008JB006257.
Fournier, M., Chamot-Rooke, N., Rodriguez, M., Huchon, P., Petit, C., Beslier, M.-O., Zaragosi, S.,
2011. Owen Fracture Zone: the Arabia–India plate boundary unveiled. Earth and Planetary Science
Letters 302, 247–252 http://dx.doi.org/10.1016/j.epsl.2010.12.027.
Gordon, R.G., DeMets, C., 1989. Present-day motion along the Owen fracture zone and Dalrymple
trough in the Arabian Sea. Journal of Geophysical Research 94, 5560-5570.
Gaedicke, C., Schlüter, H. U., Roeser, H. A., Prexl, A., Schreckenberger, B., Meyer, H., Reichert, C.,
Clift, P., Amjad, S., 2002a. Origin of the northern Indus Fan and Murray Ridge, Northern Arabian
Sea: interpretation from seismic and magnetic imaging. Tectonophysics, 355,127–143,
http://dx.doi.org/10.1016/S0040-1951(02)00137-3
Gaedicke, C., Prexl, A., Schlüter, H.-U., Roeser, H., Clift, P., 2002b. Seismic stratigraphy and
correlation of major regional unconformities in the northern Arabina Sea. In: The Tectonic and
Climatic Evolution of the Arabian Sea Region (Eds P. Clift, D. Kroon, C. Gaedicke and J. Craig),
Geological Society of London, Special Publication, 195, 25-36.
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
27
Garfunkel, Z. , Ben-Avraham, Z., 2001. Basins along the Dead Sea transform, in Peri-Tethys Memoir
6: Peri-Tethyan Rift/Wrench Basins and Passive Margins, edited by P.A. Ziegler, W. Cavazza,
A.H.F. Robertson, and S. Crasquin-Soleau, Mémoires Museum National d’Histoire Naturelle de
Paris, 186, 607–627.
Hatzfeld, D., Molnar, P., 2010. Comparisons of the kinematics and deep structures of the Zagros and
Himalaya and of the Iranian and Tibetan plateaus and geodynamic implications. Review of
Geophysics 48, RG2005, doi:10.1029/2009RG000304.
Henstock, T. J., Minshull, T. A., 2004. Localized rifting at Chagos bank in the India-Capricorn plate
boundary zone. Geology 32, 237–240.
Huang, Y., Clemens, S. C., Liu, W., Wang, L., Prell, W. L., 2007. Large-scale hydrological change
drove the Late Miocene C4 plant expansion in the Himalayan foreland and Arabian Peninsula.
Geology 35, 531-534.
Iaffaldano, G., Husson, L., Bunge , H.P., 2011. Monsoon speeds up Indian plate motion, Earth and
Planetary Science Letters 304, 503-510, doi:10.1016/j.epsl.2011.02.026.
Jolivet, L., et al., 2013. Aegean tectonics: strain localisation, slab tearing and trench retreat.
Tectonophysics 597-598, 1-33, doi:10.1016/j.tecto.2012.06.011
Kashai, E. L., Croker, P. F., 1987. Structural geometry and evolution of the Dead Sea-Jordan rift
system as deduced from new subsurface data. Tectonophysics 141, 33-60.
Kolla, V., Coumes, F., 1987. Morphology, internal structure, seismic stratigraphy, and sedimentation
of Indus Fan, The American Association of Petroleum Geologists Bulletin, pp. 650-677
Krishna, K. S., Bull, J. M., Scrutton, R. A., 2009. Early (pre-8 Ma) fault activity and temporal strain
accumulation in the central Indian Ocean. Geology 37, 227-230, doi: 10.1130/G25265A.1
Laigle, M., Hirn, A., Sachpazi, M., Roussos, N., 2000. North Aegean crustal deformation: an active
fault imaged to 10 km depth by reflection seismic data.
Lazar, M., Ben-Avraham, Z., Schattner, U., 2006. Formation of sequential basins along a strike-slip
fault – geophysical observations from the Dead Sea basin. Tectonophysics 421, 53–69.
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
28
Le Pourhiet, L., Huet, B., May, D. A., Labrousse, L., Jolivet, L., 2012. Kinematic interpretation of the
3D shapes of metamorphic core complexes. Geochemistry, Geophysics, Geosystems, 13, Q09002,
doi:10.1029/2012GC004271
Malavieille, J., 2010. Impact of erosion, sedimentation, and structural heritage on the structure and
kinematics of orogenic wedges : analog models and case studies. GSA Today 20, 4-10, doi:
10.1130/GSATG48A.1
Mann, P., 2007. Global catalogue, classification and tectonic origins of restraining- and releasing
bends on active and ancient strike-slip fault systems. Geological Society Special Publications 290,
13-142.
McCall, G.J.H., 1997. The geotectonic history of the Makran and adjacent areas of southern Iran.
Journal of Asian Earth Science 15, 517-531.
McKenzie, D. P., Sclater, J. G., 1971. The evolution of the Indian Ocean since the Late Cretaceous,
Geophysical Journal of Royal Astronomy Society 25, 437–528.
McNeill, L. C., Mille, A., Minshull, T. A., Bull, J. M., Kenyon, N. H., 2004. Extension of the North
Anatolian Fault into the North Aegean Trough: Evidence for transtension, strain partitioning, and
analogues for Sea of Marmara basin models. Tectonics 23, TC2016, doi:10.1029/2002TC001490
Merkouriev, S., DeMets, C., 2006.Constraints on Indian plate motion since 20 Ma from dense Russian
magnetic data: Implications for Indian plate dynamics. Geochemistry, Geophysics, Geosystems 7,
Q02002, doi:10.1029/2005GC001079.
Minshull, T. A., White, R. S., Barton, P. J., Collier, J. S., 1992. Deformation at plate boundaries
around the Gulf of Oman, Mar. Geol., 104, 265–277, doi:10.1016/0025-3227(92)90101-M.
Molnar, P., England, P., Martinod, J., 1993. Mantle dynamics, uplift of the Tibetan Plateau, and the
Indian monsoon. Rev. Geophys. 31, 357–396.
Molnar, P., Stock, J., 2009. Slowing of India's convergence with Eurasia since 20 Ma and its
implications for Tibetan mantle dynamics. Tectonics, 28, TC3001, doi:10.1029/2008TC002271
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
29
Morley, C. K., 2013. Discussion of tectonic models for Cenozoic strike-slip fault-affected continental
margins of mainland SE Asia. J. Asian Earth Sci., 76, 137-151,
dx.doi.org/10.1016/j.jseaes.2012.10.019
Mouchot, N., 2009. Tectonique et sédimentation sur le complexe de subduction du Makran
pakistanais. PhD thesis, 364 pp., Univ. Cergy-Pontoise, France.
Mouchot, N, Lallemant, S., Loncke, L., Leturmy, P., Mahieux, G., Chanier, F., Ellouz-Zimmermann,
N., 2008. Vertical movements and recent sedimentary processes on the Makran accretionary prism
off Pakistan, EGU General Assembly, Vienna.
Mouchot, N, Loncke, L., Mahieux, G., Bourget, J., Lallemant, S., Ellouz-Zimmermann, N., Leturmy,
P., 2010. Recent sedimentary processes along the Makran trench (Makran active margin, off
Pakistan), Marine Geology, 271, 17-31.
Mountain, G. S., Prell, W.L. 1990. A multiphase plate tectonic history of the southeast continental
margin of Oman. In: Robertson, A. H. F., Searle, M. P. and Ries, A. C. (eds) the Geology and
Tectonics of the Oman Region, Geological Society of London Special Publications, 49, 725-743.
Mouthereau, F., Lacombe, O., Vergés, J., 2012. Building the Zagros collisional orogen: Timing, strain
distribution and the dynamics of Arabia/Eurasia plate convergence. Tectonophysics 532–535, 27–
60.
Mulder, T., Ducassou, E., Gillet, H., Hanquiez, V., Tournadour, E., Combes, J., Eberli, G. P., Kindler,
P., Gonthier, E., Conesa, G., Robin, C., Sianipar, R., Reijmer, J.J.G., François, A., 2012. Canyon
morphology on a modern carbonate slope of the Bahamas : evidence of regional tectonic tilting.
Geology 40, 771-774, doi: 10.1130/G33327.1
Papanikolaou, D., Alexandri, M., Nomikou, P., Ballas, D., 2002. Morphotectonic structure of the
western part of the North Aegean Basin based on swath bathymetry. Marine Geology 190, 465-
492.
Pubellier, M., Rangin, C., Ego, F. et al. 2005. Atlas of the margins of SE Asia, N°176, 176 maps, 174
chap., 171 CDRom, Geol. Soc. France-Amer. Assoc. Petrol. Geologists.
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
30
Quittmeyer R. C., Kafka, A. L., 1984. Constraints on plate motions in southern Pakistan and the
northern Arabian Sea from the focal mechanisms of small earthquakes. Journal of Geophysical
Research 89, 2444-2458.
Reading, H.G., Richards, M., 1994. Turbidites systems in deep-water basin margins classified by grain
size and feeder system. American Association of Petroleum Geology Bulletin, 792-822.
Rodriguez, M., Fournier, M., Chamot-Rooke, N., Huchon, P., Bourget, J., Sorbier, M., Zaragosi, S.,
Rabaute, A., 2011. Neotectonics of the Owen Fracture Zone (NW Indian Ocean): structural
evolution of an oceanic strike–slip plate boundary (2011). Geochemistry, Geophysics, Geosystems
12, Q12006 http://dx.doi.org/10.1029/2011GC003731.
Rodriguez, M., Fournier, M., Chamot-Rooke, N., Huchon, P., Zaragosi, S., Rabaute, A., 2012. Mass
wasting processes along the Owen Ridge (Northwest Indian Ocean), Marine Geology 326-328, 80-
100, 10.1016/j.margeo.2012.08.008
Rodriguez, M., Chamot-Rooke, N., Hébert, H., Fournier, M., Huchon, P. 2013a. Owen Ridge deep-
water submarine landslides: Implications for tsunami hazard along the Oman coast, NHESS, 13,
417424.
Rodriguez, M., Chamot-Rooke, N., Fournier, M., Huchon, P., Delescluse, M., 2013b. Mode of
opening of an oceanic pull-apart: The 20 °N Basin along the Owen Fracture Zone (NW Indian
Ocean), Tectonics 32, 1-15, doi:10.1002/tect.20083
Rodriguez, M., Chamot-Rooke, N., Huchon, P., Fournier, M., Delescluse, M., 2014. The Owen Ridge
uplift in the Arabian Sea: implications for the sedimentary record of Indian monsoon in Late
Miocene, Earth and Planetary Science Letters 394, 1-12, doi : 10.1016/j.epsl.2014.03.011
Ross, D. A., Uchupi, E., White, R., 1986. The geology of the Persian-Gulf of Oman Region : a
synthesis. Reviews of Geophysics 24, 537-556.
Royer, J. Y., Chaubey, A. K., Dyment, J., Bhattacharya, G. C., Srinivas, K., Yateesh, V., Ramprasad,
T., 2002.Paelogene plate tectonic evolution of the Arabian and Eastern Somali basins. in The
Tectonic and Climatic Evolution of the Arabian Sea Region, edited by P. D. Clift et al., Geological
Society Special Publications 195, 7–23
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
31
Schattner, U., 2010. What triggered the early-to-mid Pleistocene tectonic transition across the entire
Eastern Mediterranean? Earth and Planetary Science Letters 289, 539-548.
Schlische, R. W., Withjack M. O., Eisenstadt, G., 2002.An experimental study of the secondary
deformation produced by oblique slip normal faulting. American Association of Petroleum
Geology Bulletin 86, 885–906.
Schlüter, H. U., Prexl, A., Gaedicke, C., Roeser, H., Reichert, C., Meyer, H., von Daniels C., 2002.The
Makranaccretionary wedge: sediment thicknesses and ages and the origin of mud volcanoes.
Marine geology 185, 219-232
Schulz, H., von Rad, U., Erlenkeuser, H., 1998. Correlation between Arabian Sea and Greenland
climate oscillations of the past 110 000 years. Nature, 393, 54-57.
Shipboard Scientific Party,1989.Site 731. In Prell, W.L., Niitsuma, N., et al., Proc. ODP, Init.Repts.,
117: College Station, TX (Ocean Drilling Program), 585-652.
Skinner, L.C. and McCave, I.N., 2003. Analysis and modelling of gravity- and piston coring based on
soil mechanics. Marine Geology 199, 181-204.
Smit J., Brun J. P, Cloetingh, S., Ben-Avraham, Z., 2008.Pull-apart basin formation and development
in narrow transform zones with application to the Dead Sea Basin. Tectonics 27, TC6018,
doi:10.1029/2007TC002119
Smit, J., Brun, J-P., Cloething, S., Ben-Avraham, Z., 2009. The rift-like structure and asymmetry of
the Dead Sea Fault. Earth and Planetary Science Letters 290, 74-82, doi:10.1016/j.epsl.2009.11.060
Smit J., Burg J-P, Dolati A., Sokoutis D., 2010. Effects of mass waste events on thrust wedges:
analogue experiments and application to the Makran accretionary wedge. Tectonics 29: TC3003
Smith, G., McNeill, L. Henstock, T. J., Bull, J., 2012. The structure and fault activity of the Makran
accretionary prism. Journal of Geophysical Research 117, B07407, doi:10.1029/2012JB009312.
Smith, G. L., McNeill, L. C., Wang, K. He, J., Henstock, T.J., 2013. Thermal structure and megathrust
seismogenic potential of the Makran subduction zone. Geophysical Research Letters 40, 1528–
1533, doi:10.1002/grl.50374.
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
32
ten Brink, U. S., Flores, C. H., 2012. Geometry and subsidence history of the Dead Sea basin: A case
for fluid induced mid-crustal shear zone? Journal of Geophysical Research 117, B01406,
doi:10.1029/2011JB008711.
Wakabayashi, J., 2007. Stepovers that migrate with respect to affected deposits: field characteristics
and speculation on some details of their evolution, in Tectonics of Strike-Slip Restraining and
Releasing Bends, edited by W.D. Cunningham and Paul Mann, Geol. Soc. Spec. Publ, 290, 169–
188.
Wang, P., Steven Clemens, S., Beaufort, L., Braconnot, P., Ganssene, G., Jiana, Z., Kershawf, P.,
Sarntheing, M., 2005. Evolution and variability of the Asian monsoon system: state of the art and
outstanding issues. Quaternary Sci. Rev. 24, 595–629.
Weissel, J. K., Anderson, R. N., Geller, C. A., 1980. Deformation of the Indo-Australian plate. Nature
287, 284-291.
Whipple, K. X., 2009. The influence of climate on the tectonic evolution of mountain belts. Nature
geoscience 2, 97-104, doi:10.1038/ngeo413
White, R. S., Klitgord, K., 1976. Sediment deformation and plate tectonics in the Gulf of Oman. Earth
and Planetary Science Letters 32, 199-209.
Whitmarsh, R.B., Weser, O. E., Ross, D. A., 1974. Initial report DSDP, U.S. Government Printing
Office, Washington, D.C., v. 23, p. 1180.
Whitmarsh, R.B., 1979. The Owen Basin off the south-east margin of Arabia and the evolution of the
Owen Fracture Zone. Geophysical Journal of the Royal Astronomical Society 58, 441-470.
Wiens, D. A., Demets, C., Gordon, R. G., Stein, S., Argus, D., Engeln, J. F., Lundgren, P., Quible, D.,
Stein, C., Weinstein, S., Woods, D. F., 1985. A diffuse plate boundary model for Indian ocean
tectonics. Geophysical Research Letters 12, 429–432.
Wu, J. E., McClay, K., Whitehouse, P., Dooley, T., 2009.4D analogue modelling of transtensional
pull-apart basins. Marine and Petroleum Geology 26, 1608-1623.
Ziegler, M., Lourens, L. J., Tuenter, E., Hilgen, F., Reichart, G-J., Weber, N, 2010. Precession
phasing offset between Indian summer monsoon and Arabian Sea productivity
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
33
linked to changes in Atlantic overturning circulation. Paleoceanography 25, PA3213
doi:10.1029/2009PA001884.
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
34
Sedimentation rates range along the Owen-Murray Ridge
Owen Ridge Murray Ridge
CORES ODP 722
(Shipboard Scientific
Party, 1989)
KS07
(Bourget et al.,
2013)
MD-2881
(Ziegker et al., 2007) S090-93KL
(Schultz et al., 1998)
Time sampled 15 Ma 160 ka 750 ka 110 ka
Average
sedimentation rates
(Plio-Pleistocene
rates)
30 m/Ma - 46 m/Ma
54 m/Ma
40 m/Ma
55 m/Ma
Table 1a
TWT (ms) Age estimates (Myr)
using P-wave velocities of 1530-1730 m.s
-1
Thickness of
pelagic
deposits
overlying the
last active
CLS -B
(OFZ-
Dalrymple
connection)
180 ± 10
3.5 ± 1.4
Thickness of
pelagic
deposits
overlying the
last active
CLS-A
(Top of
Murray Ridge)
300 ± 10
5.8 ± 2.2
Thickness of
pelagic
deposits
sealing the 'M'
unconformity
100 ± 10
1.9 ± 0.9
Table 1b
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
35
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
36
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
37
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
38
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
39
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
40
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
41
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
42
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
43
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
44
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
45
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
46
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
47
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
48
Graphical abstract
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
49
highlights
-A detailed structural framework of the Dalrymple horsetail termination of the Owen Fracture Zone
-Opening of the Dalrymple Trough around 2 Ma coeval with the regional M-unconformity
-Episode of compressive deformation along the India-Arabia plate boundary at 8-10 Ma
... Previous geoscientific studies in the region mainly focused on the relative plate motion along the OFZ between the Indian and Arabian Plates and its implications for structural patterns, stratigraphic columns, and seismicity (Mountain and Prell 1990;Rosendahl et al. 1992;Gaina et al. 2007;Fournier et al. 2008;Rodriguez et al. 2011Rodriguez et al. , 2014Rodriguez et al. , 2018among others). These investigations were primarily aimed at delineating smaller-scale features such as fault scarps, turbiditic channels, and other morphological traces retained on the seafloor, as well as sedimentary thickness, major faults, and tectonic components (ridges, seamounts, and basin). ...
... The OFZ is located north of the AOC junction, elongated further northward, and adjoins the Murray Ridge. The misnamed OFZ is Owen Transform Fault (OTF), which currently is an active dextral strike-slip boundary between Arabian-Indian plates with a differential rate of velocity (Rodriguez et al. , 2014. This differential plate motion along OFZ has been determined from geodetic and geological data (DeMets et al. 1990(DeMets et al. , 2010Fournier et al. 2008). ...
... Murray Ridge and Dalrymple Trough have a deeper Moho under a moderately thickened crust (Corfield et al. 2010). The Dalrymple Trough-Murray Ridge system experienced two deformation episodes: an initial compressional episode uplifting the southern Murray Ridge (Kolla and Coumes 1990) and a subsequent episode causing subsidence in the Dalrymple Trough and significant uplift of the northern Murray Ridge (Rodriguez et al. 2014). During the Early Miocene, the Owen-Murray Ridge (OMR) uplift led to a prominent fan shape, influencing Indus sedimentation (Rodriguez et al. , 2014(Rodriguez et al. , 2018. ...
Article
Full-text available
This study evaluates the efficacy of GECO gravity data for geophysical studies, mapping structural and tectonic features and their impact on gravity signatures in the study area. Computed correlation coefficient (96–98%), root-mean-square error (5.1–5.3 mGal), and standard deviation (3.9–4.2 mGal) between the GECO model-derived and ship-borne free-air gravity reveal the efficacy of the GECO gravity data for the geophysical studies in the region. A total horizontal derivative approach was used in order to enhance the residual and regional responses of the Bouguer gravity anomaly. The shorter-wavelength lineaments originated from subsurface mass heterogeneities were found trending in the northwest direction, subsequently east, north–northeast and east–northeast directions. In contrast, the longer-wavelength lineaments originating from deep-seated mass heterogeneities dominated in the east-northeast direction, followed by north-northeast and northwest directions. Lineaments occurring at shallower depths are associated with sedimentary/basement columns and could be utilised in basin demarcation for hydrocarbon exploration. In contrast, deep-seated lineaments originated due to deformities at the crust-mantle boundary or in the mantle and could be used in the region’s seismicity analysis. Spectral analysis and 2D forward modelling results indicate sediment thickness of ~ 2.0–6.0 km, basement thickness of ~ 6–14 km, and Moho depth of ~ 10–18 km. Delineated lineaments and computed basement and Moho depths were further validated with existing data. Anomalously high and low gravity features were interpreted based on Moho depth, basement thickness, and sediment thickness. This study concludes that anomalous gravity anomalies are mainly controlled by Moho topography despite the relatively thicker crust in the northern region. The crustal thickness mainly controls the southern latitude’s gravity signatures.
... The early drainage pattern may have been diverted to the east by the uplift of the Murray Ridge (Edwards et al., 2000), which seems to have been a multiphase process. Several growth pulses have been recognized starting in the Paleocene (Clift et al., 2001;Gaedicke et al., 2002) and peaking in the middle to late Miocene (Mountain & Prell, 1990;Rodriguez, Chamot-Rooke, Huchon, Fournier, Lallemant, et al., 2014). A sharp increase in the terrigenous influx toward the Arabian Abyssal Plain occurred in the early Miocene (Davies et al., 1995) or middle Miocene (Rea, 1992), but clays sedimentation in the northern Arabian Sea issued from the Indus River may have started earlier in the late Oligocene (Kolla & Coumes, 1987;Weedon and McCave, 1991). ...
Article
Full-text available
The Owen transform fault (OTF) connecting the Sheba and the Carlsberg spreading ridges in the Indian Ocean currently forms the active plate boundary between India and Somalia plates. This 330‐km‐long transform fault is by far the longest transform fault along the India‐Somalia plate boundary and its valley is buried under the thick distal turbidites of the Indus Fan with total thickness ranging from 1,000 to >5,000 m. A new set of seismic reflection and multibeam bathymetric data reveals remarkable transpressive structures along its entire length recorded as folds in the sedimentary cover, eruption of mud ridges at the seafloor, thrusts in the young oceanic lithosphere. Based on a new regional time‐calibration of the seismic reflectors, we show that sediments in the transform valley (post 8.6 Ma) recorded a period of tectonic quiescence until the onset of a transpressive event around 1.5–2.4 Ma that we relate to a minor change in India‐Somalia kinematics not captured by magnetic anomalies. This tectonic regime is still active based on compressive earthquakes and deformation of the most recent sediments. Transpression resulted in the formation of a proto‐median ridge and the coeval propagation of the tip of the Carlsberg Ridge into the Somalian plate. These features are typically encountered at many other transform faults but rarely captured in their very early stage.
... The azimuths of the predicted relative motion along the Arabian-Indian plate's boundary are in agreement with the right-lateral strike-slip motions prevailing over the last millions of years along the OFZ (Fournier et al., 2008;Rodriguez et al., 2019). The model predictions indicate opening motions of ∼0.7 ± 0.5 mm/yr and ∼1.3 ± 0.5 mm/ yr for the northern and southernmost extremities of the OFZ, respectively, consistent with the active normal faults associated with the Dalrymple Trough and the Beautemps-Beaupré pull-apart basin (Fournier et al., 2008;Rodriguez et al., 2014, Figures 1 and 3). The kinematic modeling results predict mild compression at ∼17.8°N, in contrast with former estimates (e.g., ArRajehi et al., 2010), but in line with the presence of restraining bends between ∼16.5°-20.3°N ...
Article
Full-text available
The present‐day motions in and around the Arabian plate involve a broad spectrum of tectonic processes including plate subduction, continental collision, seafloor spreading, intraplate magmatism, and continental transform faulting. Therefore, good constraints on the relative plate rates and directions, and on possible intraplate deformation, are crucial to assess the seismic hazard at the boundaries of the Arabian plate and areas within it. Here we combine GNSS‐derived velocities from 168 stations located on the Arabian plate with a regional kinematic block model to provide updated estimates of the present‐day motion and internal deformation of the plate. A single Euler pole at 50.93 ± 0.15°N, 353.91 ± 0.25°E with a rotation rate of 0.524 ± 0.001°/Ma explains well almost all the GNSS station velocities relative to the ITRF14 reference frame, confirming the large‐scale rigidity of the plate. Internal strain rates at the plate‐wide scale (∼0.4 nanostrain/yr) fall within the limits for stable plate interiors, indicating that differential motions are compensated for internally, which further supports the coherent rigid motion of the Arabian plate at present. At a smaller scale, however, we identified several areas within the plate that accommodate strain rates of up to ∼8 nanostrain/yr. Anthropogenic activity and possible subsurface magmatic activity near the western margin of the Arabian plate are likely responsible for the observed local internal deformation. Put together, our results show a remarkable level of stability for the Arabian lithosphere, which can withstand the long‐term load forces associated with active continental collision in the northeast and breakup to the southwest with minimal internal deformation.
... In the Oman abyssal plain, the top of U4 is in continuity with the top Oligocene-Early Miocene unconformity recognized on the western flank of the Qalhat Seamount (Edwards et al., 2000;Rodriguez et al., 2016). This somewhat diachronous unconformity is observed across the whole Indian Ocean (Gaedicke et al., 2002;Rodriguez et al., 2014). ...
Article
The northern Oman margin is a key area for understanding the emplacement of the Semail Ophiolite and obduction processes in general. This study used a grid of 2D-multichannel seismic lines linked to well data to characterize the offshore domain of the Semail Ophiolite and reappraised the obduction and post-obduction history of the Oman margin. West of Muscat, in the Sohar basin, the Late Cretaceous to Paleogene tectonic mega-sequence records syn- to late-obduction stages and the deposition of erosional products of the autochthonous Arabian sediments, including a major mass transport complex. Syn-obduction thrusting is documented in this sector only, as a major fault emplacing a distal basement high (likely volcanic) onto Campanian sediments over >10 km. To the east, the Hatat and Tiwi basins are characterized by a less copious Maastrichtian-Paleogene sequence. These basins developed above a domain characterized by the northern equivalent of the Saih Hatat Dome and later extensional faults. This sector distinctively records the extensional phase associated with the exhumation and erosion of the subducted continental margin. The dichotomy between the two sectors is linked due to a structural high located offshore, in the continuation of the Semail Gap transfer fault. We propose that this transfer fault, coincident with a major Pan-African structure, affected the architecture of the passive margin during both rifting in the Neotethys and later ophiolite emplacement, i.e. during (continental) subduction and obduction.
... In the Oman abyssal plain, the top of U4 is in continuity with the top Oligocene-Early Miocene unconformity recognized on the western flank of the Qalhat Seamount (Edwards et al., 2000;Rodriguez et al., 2016). This somewhat diachronous unconformity is observed across the whole Indian Ocean (Gaedicke et al., 2002;Rodriguez et al., 2014). ...
Article
Full-text available
The northern Oman margin is a key area for understanding the emplacement of the Semail Ophiolite and obduction processes in general. This study uses a grid of 2D‐multichannel seismic lines tied to well data to characterize the offshore domain of the Semail Ophiolite and reappraises the obduction and post‐obduction history of the Oman margin. West of Muscat, in the Sohar basin, the late Cretaceous to Paleogene tectonic mega‐sequence records syn‐ to late‐obduction stages and the deposition of erosional products of the autochthonous Arabian sediments, including a major mass transport complex. Syn‐obduction thrusting is documented in this sector only, as a major fault emplacing a distal basement high (likely volcanic) onto Campanian sediments over >10 km. To the east, the Hatat and Tiwi basins are characterized by a less‐copious Maastrichtian‐Paleogene sequence. These basins developed above a domain characterized by the northern equivalent of the Saih Hatat dome and later extensional faults. This sector distinctively records the extensional phase associated with the exhumation and erosion of the subducted continental margin. The dichotomy between the two sectors is linked due to a structural high located offshore, in the continuation of the Semail Gap transfer fault. We propose that this transfer fault, coincident with a major Pan‐African structure, affected the architecture of the passive margin during both rifting of the Neotethys and later ophiolite emplacement, that is, during (continental) subduction and obduction.
... It is directed NW until the 10°N. At 10°N, the Carlsberg Ridge is dissected by the large transverse Owen Fracture Zone with a total length in the adjacent basins of 2,800 km, including Chain Ridge and Murray Ridge (Rodriguez et al. 2014). Then, it changes direction to almost 90° until 12.5°N before continuing as the Aden-Sheba Ridge in the Gulf of Aden (Fig. 2). ...
Article
Full-text available
This study examined the relationships between topographic structure and submarine geomorphology, sediment thickness, geophysical anomaly fields, geological settings and tectonic lineament stretching of the Arabian Sea region, Carlsberg Ridge morphology, Makran Trench depths by GMT. The study included spatial analysis of the high-resolution datasets (GEBCO, EGM96, GlobSed) and geomorphological modeling of the 300km-width cross-section profiles of the Makran Trench. The analysis shown correlation between complex geologic and tectonic structure, asymmetric geomorphology and geophysical anomaly fields. The Makran Trench is formed in the subduction zone of the Arabian and Eurasian plates at the basement of the continental margin of Pakistan. Submarine geomorphic structure of the Arabian Sea is complicated by the Carlsberg Ridge, Owen Fracture Zone, Aden-Owen- Carlsberg Triple junction, numerous faults and rifts. The geophysical fields of the marine free- air gravity correlate with distribution of these geomorphic structures. Bathymetric analysis of the trench revealed the most frequent depth (448 samples) at -3,250 to -3,500 m, following by intervals: -3,000 to -3,250 m (225 samples), -2,750 to -3,000 m (201 samples). Gently declining continental slope of the coastal elevations correlate with gradually decreasing depths, as equally distributed bins: 124 samples (-2,500 to -2,750 m), 96 (-2,250 to -2,500 m), 86 (-2,000 to -2,250 m). Makran Trench has asymmetric geomorphology with a high slope steepness on the continental slope of Pakistan and low steepness with flat valley on the oceanward side.
Conference Paper
Full-text available
Pakistan is endowed with two major sedimentary basins, namely Balochistan and Indus, separated by left-lateral Bela-Ornach-Chaman transform fault system in the onshore and Murray Ridge in the offshore. Both basins are different in context geological evolution, extending from land to Arabian Sea Shelf. NE-SW-trending right-lateral strike-slip fault of Murray Ridge dividing Makran Offshore basin in the west and the Indus Offshore basin in the east. Indus Basin is proven petroleum province. Whereas Makran Complex is still virgin to flow hydrocarbons. So far 18 well has been drilled in Pakistan Offshore without any commercial discovery. Out of these wells 3 were drilled in Makran Offshore and 15 were drilled in Indus Offshore. This paper presents an over review of different tectonic elements as shown in Fig.-1, basin evolution, exploration history, possible hydrocarbon traps based on seismic data and petroleum geology of Pakistan Offshore. Brief discussion on the possibilities of gas hydrates in Makran Offshore will also be discussed.
Presentation
Full-text available
Pakistan is endowed with two major sedimentary basins, namely Balochistan and Indus, separated by left-lateral Bela-Ornach-Chaman transform fault system in the onshore and Murray Ridge in the offshore. Both basins are different in context geological evolution, extending from land to Arabian Sea Shelf. NE-SW-trending right-lateral strike-slip fault of Murray Ridge dividing Makran Offshore basin in the west and the Indus Offshore basin in the east. Indus Basin is proven petroleum province. Whereas Makran Complex is still virgin to flow hydrocarbons. So far 18 well has been drilled in Pakistan Offshore without any commercial discovery. Out of these wells 3 were drilled in Makran Offshore and 15 were drilled in Indus Offshore. This paper presents an over review of different tectonic elements as shown in Fig.-1, basin evolution, exploration history, possible hydrocarbon traps based on seismic data and petroleum geology of Pakistan Offshore. Brief discussion on the possibilities of gas hydrates in Makran Offshore will also be discussed.
Article
Full-text available
The dynamics of slab detachment and associated geological fingerprints have been inferred from various numerical and analogue models. These invariably use a setup with slab-pull-driven convergence in which a slab detaches below a mantle-stationary trench after the arrest of plate convergence due to arrival of continental lithosphere. In contrast, geological reconstructions show that post-detachment plate convergence is common and that trenches and sutures are rarely mantle-stationary during detachment. Here, we identify the more realistic kinematic context of slab detachment using the example of the India-Asia convergent system. We first show that only the India and Himalayas slabs (from India’s northern margin) and the Carlsberg slab (from the western margin) unequivocally detached from Indian lithosphere. Several other slabs below the Indian Ocean do not require a Neotethyan origin and may be of Mesotethys and Paleotethys origin. Additionally, the still-connected slabs are being dragged together with the Indian plate forward (Hindu Kush) or sideways (Burma, Chaman) through the mantle. We show that Indian slab detachment occurred at moving trenches during ongoing plate convergence, providing more realistic geodynamic conditions for use in future numerical and analogue experiments. We identify that the actively detaching Hindu Kush slab is a type-example of this setting, whilst a 25-13 Ma phase of shallow detachment of the Himalayas slab, here reconstructed from plate kinematics and tomography, agrees well with independent, published geological estimates from the Himalayas orogen of slab detachment. The Sulaiman Ranges of Pakistan may hold the geological signatures of detachment of the laterally dragged Carlsberg slab.
Article
Full-text available
Regional seismic data with deep wells information has been used to interpret the depositional setting of Miocene sediments in the Offshore Indus Basin, Pakistan. The current study has implications for understanding the complex subsurface geology of the Offshore Indus Basin, Pakistan. Two seismic profiles were selected to interpret the sedimentary packages within the Miocene sediments. The interpreted seismic data showed four different recognizable system tracts in the Miocene sediments. These include Lowstand, Highstand and Falling Stage System Tracts. The sedimentary packages differ in regional setup depending on the location within the offshore basin. The eastern part (as highlighted by seismic Profile-2 and Profile-3) exhibits more complex sedimentary geometries in comparison to western part (as highlighted by seismic Profile-1). The earlier deposition is controlled by lowstand settings when the sea-level dropped below the shelf margin leading to incision over the shelf. This led to the deposition of Lowstand fan systems. This was followed by a rise in sea-level leading the way for the carbonate deposition over the shelf. An intermittent unconformity is identified by another set of lowstand fans and truncation patterns. This unconformity halted the carbonate deposition over the shelf and transported the sediments to the basin floor. Later, rise in sea-level resumed the carbonate deposition over the shelf. Another sea-level drop is noticed and sediments are interpreted to be moving towards the basin leading to a Falling Stage System Tract and easily identified by regressive onlaps. This unconformity is more pronounced towards east where it eroded the carbonates. Downlaps seen above this unconformity confirm the deepening conditions leading the way for the deposition of finer sediments. Further drop in sea-level is seen leading the way for coarse sediment deposition. We believe that this study would provide strong support for researching global passive margins from economic perspective as well as understanding complex depositional framework.
Article
Full-text available
The upper Indus Fan (1,600-3,600 m) is characterized by up to several hundred meters relief that resulted from the aggradation of large channel-levee complexes; gradients greater than 1:500; a distinct 3.5 kHz echo character with several continuous subbottom reflectors; and by fine-grained sediments, except within the channels where coarse-grained materials are inferred. The lower fan (4,000-4,500 + m) has a smooth relief with channels and levees of relatively small dimensions; overall gradients of less than 1:1,000; prolonged 3.5 kHz echo character with few or no subbottom reflectors; and a dominantly sandy lithology. The characteristics of the middle fan are intermediate between those of the upper and lower fans. Seismic records reveal at least three canyon complexes on the shelf, each of which gave rise to several leveed channels on the fan. The canyons and channels migrated extensively in time and space across the fan, and channel abandonment and avulsion were very common. Seismically, the canyon fill consists of several reflection-free zones overlain by inclined reflections of moderate amplitudes which are inferred to indicate fine-grained sediments. The channel fills consist of high-amplitude, random reflections overlain successively by reflection-free zones and weak to moderate-amplitude continuous reflections. These characteristics suggest coarse-grained deposits at the base fining upward to the top of the channel fill. The channels, especially on the upper and middle fan, are flanked by wedge-shaped, concave-upward reflection packages characteristic of levee-overbank deposits. Sea level changes and the Himalayan orogenies have profoundly affected the Indus Fan sedimentation since the Oligocene-early Miocene. Sedimentation was dominantly by channelized turbidity currents with overbank deposition on the upper fan, and by both unchannelized and channelized turbidity currents on the lower fan during the lowstands of sea level.
Article
Full-text available
New high-quality multibeam data presented here depict the northern slope of the Little Bahama Bank (Bahamas). The survey reveals the details of large- and small-scale morphologies that look like siliciclastic systems at a smaller scale, including large-scale slope failure scars and canyon morphologies, previously interpreted as gullies and creep lobes. The slope exhibits mature turbidite systems built by mass-flow events and turbidity currents. The sediment transport processes are probably more complex than expected. Slope failures show sinuous head scarps with various sizes, and most of the scars are filled with recent sediment. Canyons have amphitheater-shaped heads resulting from coalescing slump scars, and are floored by terraces that are interpreted as slump deposits. Canyons rapidly open on a short channel and a depositional fan-shaped lobe. The entire system extends for similar to 40 km. The development of these small turbidite systems, similar to siliciclastic systems, is due to the lack of cementation related to alongshore current energy forcing the transport of fine particles and flow differentiation. Detailed analyses of bathymetric data show that the canyon and failure-scar morphology and geometry vary following a west-east trend along the bank slope. The changing parameters are canyon length and width, depth of incision, and canyon and channel sinuosity. Accordingly, failure scars are larger and deeper eastward. These observations are consistent with a westward tectonic tilt of the bank during the Cenozoic.
Article
Full-text available
Australia and India are conventionally thought to be contained in a single plate divided from an Arabian plate by the Owen Fracture Zone. We propose instead that motion along the nearly aseismic Owen Fracture Zone is negligible and that Arabia and India are contained within a single Indo-Arabian plate, divided from the Australian plate by a diffuse boundary. This boundary, which trends E-W from the Central Indian Ridge near Chagos Bank to the Ninetyeast Ridge, and north along the Ninetyeast Ridge to the Sumatra Trench, is a zone of concentrated seismicity and deformation heretofore characterized as “intraplate”. Plate motion inversions and an F-ratio test show that relative motion data along the Carlsberg Ridge are fit significantly better by the new model. The misclosure of the Indian Ocean triple junction is reduced by 40%. The rotation vector of Australia relative to Indo-Arabia is consistent with the seismologically observed ∼2 cm/yr of left-lateral strike-slip along the Ninetyeast Ridge, N-S compression in the Central Indian Ocean, and the N-S extension near Chagos. This boundary, possibly initiated in late Miocene time, may be related to the opening of the Gulf of Aden and the uplift of the Himalayas. The convergent segment of this boundary may represent an early stage of convergence preceding the initiation of subduction.
Article
Depositional systems in deep-water basin margins can be classified on the basis of grain size and feeder system into 12 classes: mud-rich, mud/sand-rich, sand-rich, and gravel-rich "point-source submarine fans;" mud-rich, mud/sand-rich, sand-rich, and gravel-rich "multiple-source submarine ramps;" and mud-rich, mud/sand-rich, sand-rich, and gravel-rich "linear-source slope aprons." The size and stability of channels and the organization of the depositional sequences decreases toward a linear source as does the length:width ratio of the system. As grain size increases, so does slope gradient, impersistence of channel systems, and tendency for channels to migrate. As grain size diminishes, there is an increase in the size of the source area, the size of the depositional sys em, the downcurrent length, the persistence and size of flows, fan channels, channel-levee systems, and in the tendency to meander and for major slumps and sheet sands to reach the lower fan and basin plain. The exact positioning of any one depositional system within the scheme cannot always be precise, and the position may be altered by changes in tectonics, climate, supply, and sea level. However, the models derived from each system are sufficiently different to significantly affect the nature of petroleum prospectivity and reservoir pattern. Understanding and recognizing this variability is crucial to all elements of the exploration-production chain. In exploration, initial evaluations of prospectivity and commerciality rely on the accurate stratigraphic prediction of reservoir facies architecture, and trapping styles. For field appraisal and reservoir development, a similar appreciation of variability aids reservoir description by capturing the distribution and architecture of reservoir and nonreservoir facies and their impact on reservoir delineation, reservoir behavior, and production performance.
Article
The sedimentary construction of oceanic margins is most often carried out by the combined action of gravitational processes and processes related to bottom (contour) currents. One of the major difficulties encountered in the interpretation of seismic profiles crossing such margins is the differentiation of these two types of deposit, especially where they display very complicated imbricated geometries. The aim of this paper, therefore, is to derive criteria for the recognition of contourite vs. turbidite deposits, based on the analysis of many seismic profiles from both published and unpublished sources. The following features are the most diagnostic for the recognition of contourite drifts. At the scale of the basin, four different drift types can be distinguished according to the morphostructural context, their general morphology and the hydrodynamic conditions. These are: contourite-sheeted drifts (including abyssal sheets and slope-plastered sheets), elongate-mounded drifts (detached and separated types), channel-related drifts (including lateral and axial patch drifts and downstream contourite fans), and confined drifts trapped in small, tectonically active basins. At the scale of the drift, three features provide the best diagnostic criteria for recognising contourite deposits on seismic profiles: major discontinuities that can be traced across the whole drift and represent time lines corresponding to hydrological events, lenticular, convex-upward depositional units with a variable geometry, and a specific style of progradation–aggradation of these units that is influenced by interaction of the bottom current with Coriolis force and with the morphology. At the scale of depositional units, the seismofacies show a wide variety of reflector characteristics, many of which are very similar to those observed in turbidite series. Distinction between sediment wave seismofacies deposited by turbidity currents and bottom currents still remains ambiguous.
Article
The Corinth rift (Greece) is one of the world's most active rifts. The early Plio‐Pleistocene rift is preserved in the northern Peloponnese peninsula, south of the active Corinth rift. Although chronostratigraphic resolution is limited, new structural, stratigraphic and sedimentological data for an area >400 km2 record early rift evolution in three phases separated by distinct episodes of extension rate acceleration and northward fault migration associated with major erosion. Minimum total N–S extension is estimated at 6.4–7.7 km. The earliest asymmetrical, broad rift accommodated slow extension (0.6–1 mm a−1) over >3 Myrs and closed to the west. North‐dipping faults with throws of 1000–2200 m defined narrow blocks (4–7 km) with little footwall relief. A N‐NE flowing antecedent river system infilled significant inherited relief (Lower group). In the earliest Pleistocene, significant fluvial incision coincided with a 15 km northward rift margin migration. Extension rates increased to 2–2.5 mm a−1. The antecedent rivers then built giant Gilbert‐type fan deltas (Middle group) north into a deepening lacustrine/marine basin. N‐dipping, basin margin faults accommodated throws −1. This transition may correspond to an unconformity in offshore lithostratigraphy. Middle group deltas were uplifted and incised as new hangingwall deltas built into the Gulf (Upper group). A final increase to present‐day extension rates (11–16 mm a−1) probably occurred in the Holocene. Fault and fault block dimensions did not change significantly with time suggesting control by crustal rheological layering. Extension rate acceleration may be due to strain softening or to regional tectonic factors.