Geological Society, London, Special Publications: Geological Development of The Timor Orogen
Geological Society, London, Special Publications: Geological Development of The Timor Orogen
Geological Society, London, Special Publications: Geological Development of The Timor Orogen
Accepted Manuscript
© 2019 The Author(s). Published by The Geological Society of London. All rights reserved. For
permissions: http://www.geolsoc.org.uk/permissions. Publishing disclaimer:
www.geolsoc.org.uk/pub_ethics
Although reasonable efforts have been made to obtain all necessary permissions from third parties to include their
copyrighted content within this article, their full citation and copyright line may not be present in this Accepted Manuscript
version. Before using any content from this article, please refer to the Version of Record once published for full citation and
copyright details, as permissions may be required.
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
PT
CCG Multi-Client & New Ventures, Crawley, UK
RI
PMD https://orcid.org/0000-0002-2253-9336
*Correspondence ([email protected])
SC
Abstract: The Timor Orogen comprises the island of Timor, a narrow offshore area to the north and a wider offshore fold-
and-thrust belt to the south. This orogen formed by jamming and subsequent collision of the Banda Sea subduction system
by the Australian plate. The BandaSeis seismic survey has revealed excellent images of the deepwater fold-and-thrust belt.
U
Seismic interpretation of the dataset demonstrated structural and tectonic features not previously described, including
N
regional geological features on the Australian continental crust and two regional NE-SW sinistral strike-slip faults and a
prominent Middle Permian paleogeographic high (Timor Plateau). Moreover, since the Middle–Late Triassic and Middle
A
Jurassic, the two NE-trending strike-slip faults governed the formation of the West Timor and Cova-Lima sub-basins. The
M
location along the Australian margin plays a dominant role controlling the structural style and shaping of the Timor Orogen.
Vertical loading and the southerly motion of the orogenic wedge are the main driving forces responsible for its building,
ED
illustrating a thin-skinned tectonic framework. Thrust faults nucleate in a forward-breaking sequence in the motion of
thrust transport with younger thrusts developing in front of older thrusts. Most of the collisional deformation has been
classified into two styles: shallow thin-skinned and deep-seated deformation.
T
The horseshoe-shaped Banda Sea is arguably the most eye-catching feature of South East Asia (SEA)
EP
geography (Fig. 1). This region is the product of complex collision among the Eurasia, Australia, and
Pacific (Caroline and Philippine Sea) tectonic plates, ongoing since the Late Oligocene (many
CC
references, but see (Hamilton 1979; Hall 2002, 2012; Hinschberger et al. 2005) (Fig. 2). Numerous
studies have been undertaken on various aspects of the geology of the region, but many questions
remain unresolved. As noted by Bird (2003), the region “has the most complicated neotectonics on
A
Earth, with exceptionally high rates of vertical movements and rapid but variable lateral motions that
are not easily described in terms of plate tectonics”.
Neogene geodynamics of these four plates were partially constrained by the NUVEL-1A plates model
(DeMets et al. 1990, 1994). Later on, SEA was fully constrained with regional-focused models with
Cenozoic reconstructions (Hall 2002), and lately extended to Cretaceous times (Zahirovic et al.,
2014). Finally, newer models such as MORVEL plates (DeMets et al., 2010), and lately the GSRM v.2.1
plate model cover the SEA region (Kreemer et al., 2014). With a different approach, Seton et al. 2012
proposed a model considering plate boundaries as dynamically evolving features. Recently, a model
linking surface plate motions with subducted slab remnants mapped from seismic tomography have
been proposed (Williams et al., 2015) using subduction histories (Seton et al., 2012).
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
Moreover, present-day motions are well known from GPS measurements (Genrich et al., 1996;
Kreemer et al., 2000, 2014; Bock et al., 2003; Nugroho et al., 2009; Koulali et al., 2017) as well as slip
vector studies (McCaffrey 1992, 1996) and strain rates in plate boundary zones (Kreemer et al.,
2014).
Geological background
Detailed discussion of the subsequent development of the Meso-Tethys and Ceno-Tethys ocean
realms (Metcalfe, 2009; Hall, 2012) is beyond the scope of this paper but suffice to note that from at
PT
least 45 Ma until ~9.8–5 Ma, oceanic and transitional crust of the Indo-Australian plate has been
obliquely subducted towards the NNE at ~63–75 mm/yr at the Sunda trench (Fig. 2). This subduction
process formed the Sunda Arc, extended from Sumatra eastward to Flores Island (Hartono, 1990;
RI
DeMets et al., 1994; Hall, 2002; Bock et al., 2003; Müller et al., 2008).
SC
The distinctive horseshoe-shaped of the Banda Arc is the result of a intricate relationship of
inherited Jurassic and Palaeozoic structures coupled with the oblique subduction of the Australian
plate within a complex collisional framework among the Eurasia, Australia, and the Pacific plates
U
(Figs 1, 2). The current morphology of the front of subduction of the Banda arc has been associated
to the Late Jurassic rifting and breakup of part of the East Gondwana margin, which resulted in the
N
formation of an embayed continental shelf, dubbed as Banda Embayment, of newly-formed Jurassic
A
ocean crust (Longley et al., 2002; Keep et al., 2003; Hall, 2012) (Fig. 2).
M
The Banda Arc system, open to the west, is globally unique, and has triggered robust arguments
about whether it reflects a single, strongly folded subducting slab, or two slabs; one dipping south
ED
below Seram, and the Australia Plate dipping northward below the south arm of the Banda Arc (Fig.
1).
Immediately south of the Banda Sea, the gently-curved Timor Orogen comprises the island of Timor,
T
a narrow offshore zone to the north and a wider offshore area to the south, terminating at the Timor
EP
Trough. Timor Island is part of the topographic expression of the Timor Orogen seated where the
Australian continental crust still dips very gently northwardly before sinking down towards the
Banda Sea. The western extent includes the island of Rote and the offshore area to the southeast. Its
CC
eastern extent is not well defined, but includes the Leti islands and terminates (arbitrarily) near the
island of Babar. The southern offshore area comprises the fold-and-thrust belt discussed here (Fig.
A
1).
The island of Timor comprises juxtaposed units of different tectonostratigraphic affinities. The
predominance of shales in the stratigraphy commonly leads to the formation of tectonic melange
and broken ground in place of coherent thrust sheets (Harris et al., 1998; Charlton, 2002; Haig et al.,
2008) (Fig. 3). Timor is a rare modern example of arc-continent collision and subsequent accretion:
the onset of the orogenic wedge was triggered because subduction was jammed as thicker
continental crust arrived at the subduction zone (Haig, 2012).
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
The collision between Australian continental crust and the Banda embayment (also named as Argo-
Tanimbar-Seram Ocean) began at ~25 Ma in the New Guinea region (collision with the Pacific and
Caroline plates) and continued diachronously westwards (Hall, 2002; Keep et al., 2003) (Fig. 2).
The collision in Timor Island sector, located in the Outer Banda Arc, led to hampering and
progressively locking up the subduction system east of Flores Island (Fig. 2).
The timing of the arrival of the Australian continental crust has been widely reviewed and discussed
by Audley-Charles et al. (1988), Hartono (1990), Sani et al. (1995), McCaffrey (1996), Richardson &
Blundell (1996), Audley-Charles (2004, 2011), Spakman & Hall (2010), Keep & Haig (2010), Hall
PT
(2011) and Haig (2012). In summary, there are two different schools of thought, varying from ~10–7
Ma for some researchers to >4–2 Ma for others.
RI
Thorough stratigraphic findings (Haig & McCartain, 2007; Keep & Haig, 2010; Haig, 2012) and
structural restorations (Tate et al., 2015) suggest that Australian continental crust entered into the
SC
subduction zone in the Timor region during the Late Miocene, between ~9.8 and 5.7 Ma, causing
disruption and eventual jamming of the subduction zone at approximately 5.7 Ma. Consequently,
collision caused subduction to cease, evidenced by a present-day lack of deep seismic activity and
U
cessation of volcanism in the arc immediately north of Timor (Chamalaun & Grady, 1978; Ely et al.,
2011).
N
A
This age for collision, after 9.8 Ma but before 5.7 Ma (Haig & McCartain, 2007; Roosmawati & Harris,
2009; Keep & Haig, 2010; Haig, 2012; Tate et al., 2015) is older than some other authors estimates
M
between 4 to 2 Ma (Sani et al., 1995; Richardson & Blundell, 1996; Audley-Charles, 2004; Spakman &
Hall, 2010; Ely et al., 2011; Hall et al., 2011).
ED
This earlier time window is as a result of structural restorations from field studies (Kolbano Block)
and geochronological studies from onshore data. In particular, this younger estimation is supported
by 40Ar/39Ar geochronology of bi-modal subaqueous volcanism recorded in the Ataúro volcano (see
T
These findings confirm that volcanism and subsequent subduction of the Australian lithosphere
continued until near 3.3 Ma (Ely et al., 2011). However, two facts make the Ataúro lavas a less
CC
accurate approach to dating the onset of collision: the obliquity convergence and the array of dacite
to rhyolite compositions indicating crustal contamination. This crustal contamination has important
implications to consider such as arc magma depth generation, nature of the Australian crust and
A
convergence velocities.
In order to reach the depth of arc magma generation constrained by several authors between 86–
162 km (Gill, 1981; Stern, 2002), 65–130 km (England et al., 2004) or a wider range between 72–173
km (Syracuse & Abers, 2006), with the convergence velocities between ~63–75 mm/yr (DeMets et
al., 1994; Hall, 2002; Bock et al., 2003; Müller et al., 2008), the subducted Australian continental
crust would have required at least 1 m.y. considering the fastest and shallowest scenario and more
than 2.5 m.y. in the deepest condition with a convergence velocity of 63 mm/yr. Therefore, the
crustal contamination of the Ataúro subaqueous volcanism would be sourced by an Australian
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
continental crust which would have reached the subduction trench between 4.3 to 5.8 Ma ago and
between 1–2.5 m.y. later, reaching the optimum depth for magma generation at 3.3 Ma (Fig. 1).
In regard to the collision between Australian continental crust and Sundaland, the shape of the
Jurassic embayment and age of the oceanic lithosphere within the Australian plate were major
influences on the way the collision subsequently developed (Hall, 2011). Hence, the morphology of
the subducting plate is a paramount feature to consider (Fig. 2). The succession of several
extensional events such as Pillara (Middle–Late Devonian), Westralian (Late Carboniferous–Early
Permian) and finally the Latest Callovian, most probably would have created a hyper-extended
continental margin in the North West Shelf (Heine & Müller, 2005; Heine et al., 2012; Pryer et al.,
PT
2014; Belgarde et al., 2015). Therefore, it is reasoned that the arrival of such attenuated crust at the
Timor trench will show a mechanical/tectonic response little different from typical oceanic crust
until either progressively thicker continental crust slides into the subduction channel, or the
RI
buoyancy effects of accumulated subducted continental crust leads to gradual choking of the
subduction zone, regional uplift, and the development of fold-and-thrust belts and linked foredeeps
SC
such as the Timor Trough.
This scenario has been well imaged by ambient noise tomography of the volcanic arc under Flores,
U
Wetar, and Alor (Fig. 1). Tomography images show a clear low velocity anomaly at the transition
N
between subduction of oceanic and continental lithosphere (Porritt et al., 2016). Furthermore, this
proposition is widely supported by geochemistry surveys, especially isotope analyses (helium, lead,
A
molybdenum, neodymium, strontium and oxygen) indicating the Australian continental crust as part
of the source of arc magmas (Wheller et al., 1987; Hilton & Craig, 1989; Hilton et al., 1992; Vroon et
M
al., 1995; Herrington et al., 2011; Wille et al., 2018). These findings are in line with the geochemical
dacite to rhyolite compositions indicating crustal contamination in the Ataúro lavas (Fig. 1).
ED
Remarkably, Vroon et al. (1995) found isotopic variations with an increase in 206Pb/204Pb ratio and a
decrease in the 143Nd/144Nd ratio along northeast to southwest along the Banda Arc. Moreover,
T
these lead isotopic variations in the shelf and wedge sediments along the Banda Arc are parallel to
similar variations in volcanic materials, being a strong evidence of the incorporation of subducted
EP
Furthermore, Lu/Hf, Zr/Hf and Hf-Nd isotope systematics of the Banda Arc have demonstrated the
presence of subducted continental material and the amount of continental signature increases along
the arc from <2% in the NE to >2% in the SW (Nebel et al., 2011). This data suggests that transitional
A
continental crust begun to be subducted earlier in Timor Island than Banda Sea, so that likely,
collision was earlier in Timor (Fig. 1).
Consequently, it is concluded that the collision was not a coeval event along-strike of the Timor
Trough, as suggested by Roosmawati & Harris (2009) based on analysis of foraminifera of
synorogenic pelagic units in Rote and Savu islands.
In the south arm of Banda Arc two features underpin the diachronous collisional event: the
Australian-Sundaland oblique relationship and the irregular shape of the Australian continental
margin with promontories such as the Scott Plateau and oceanic embayment such as the Argo-
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
Tanimbar-Seram oceanic gulf. Accordingly, the age of collision initiation should vary along this
rugged ancient palaeotopography (Figs. 1-2).
Hilton & Craig (1989) proposed that helium isotope variations on lavas could be linked with NW-SE
transcurrent faulting (see also Eva et al., 1988) as regional features to control and place continental
material underneath the Savu Basin immediately south of Flores Island. These transcurrent faults
would be those compartmentalising the inherited palaeogeograhy of the Australian continental
irregular edge that collided with Sundaland (Fig. 2).
Thus, it seems plausible to consider the timing of the collision in Timor Island as a diachronous
PT
event, being older in Timor-Leste and younger to the West Timor in Kolbano Block and Banda Sea.
Hence, those foredeep synorogenic packages would likely record the latest stages of the oblique
collision and its irregular shape arriving along-strike to the subduction trench in the Banda Arc
RI
system.
SC
BandaSeis seismic dataset
U
Broadband seismic data offers valuable new insights into the geology offshore Timor and an
outstanding tool to reconcile these differing tectonic interpretations. The primary purpose of this
N
study is to match the findings from seismic interpretation with the different hypotheses or theories
A
proposed for the different issues in the Timor Trough. In addition, we consider the role played by the
palaeogeography of the Australian continental margin in shaping the outer wedge of the Timor
M
Orogen.
The tectono-stratigraphy in the study area is represented by the late Carboniferous to Middle
ED
Jurassic Gondwana Megasequence deposited within interior basins of the East Gondwana rift system
(Harris et al., 1998; Harris, 2006; Haig et al., 2007; Haig & McCartain, 2010), Late Jurassic to early
Late Miocene Australian Margin Megasequence deposited on the passive margin and the
T
synorogenic succession (Haig et al., 2007; Roosmawati & Harris, 2009; Haig, 2012) deposited in the
EP
The study area covered by the Broadband BandaSeis 2D seismic survey is located offshore along with
CC
the south coast of Timor Island following the ENE-trending trench created by the collision between
Sundaland and Australia plates (Figs 1, 2). The survey was acquired during 2013–2015 in five phases
using CGG broadband technologies combining the broadest bandwidth (6 octaves) and sourced by
A
airgun with a shot interval of 25m, a seismic receiver trace interval of 12.5m and a sampling rate of
2ms.
The seismic data is clearly interpretable down to middle Permian reflectors which define the
morphology of the Palaeozoic structures (Figs 3, 4).All wells used to identify the age of the different
horizons picked in the seismic data are in the south edge of the seismic data (Fig. 5). In the western
portion, those wells are: Belalang-1, Manta-1, Mina-1 and Napoleon-1. In the central and eastern
areas of the survey the wells used to constrain the seismic data were Cova-1 and Sera TLK-A1 (Fig.
5).
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
PT
results in morphological elevation of the outer ridge. In general, the more material scraped off, the
higher the elevation of the wedge (Fig. 6). The underplating process resembles large-scale thrust
sheets in mountain ranges. Previously juxtaposed layers of sediment are stacked during the
RI
shortening process, becoming more and more vertical as younger thrusts are accreted in the wedge-
toe front, pushing older thrusts backward (Figs 4, 6, 7, 8). Therefore, older sediments are emplaced
SC
over younger rocks.
The stress field changes dramatically from one sector to the other. Whereas σ1 is vertical in the
U
Australian Plate creating a reactivation of normal faults, this σ1 vector shifts to horizontal due to the
collision between the underplating plate and the wedge (Langhi et al., 2011). The structural pattern
N
of the downward plate in the study area can be divided into two sectors, western and eastern, with
A
each showing different fault population and structural geometries (Fig. 9).
M
In summary, both sectors show a regional NE-trending sinistral transtensional strike-slip faults with a
synformal sub-basin in its hanging-wall and a pronounced relief, its footwall being on the eastern
side. The western and eastern sub-basins are the West Timor and Cova Lima sub-basins (Fig. 9).
ED
The NE-trending Nunkolo sinistral strike-slip fault shows a throw of more than 3,600m dipping to the
EP
WNW. This fault outlines the west flank of a 4-way closure in the Late Carboniferous horizon and
mirrored to the Jurassic horizons. Hence, though the seismic data does not image the basement it is
quite reasonable to assume that this regional fault could be a basement-involved fault. In the WNW
CC
flank of this regional fault a negative flower structure can be described with all antithetic faults
rooted in the Nunkolo strike-slip fault. According to the seismic interpretation, the Nunkolo fault
opened the West Timor sub-basin, opposite to the Indonesian province of West Timor, during the
A
The West Timor Sub-basin shows an elongated NE-trending axis parallel to the Nunkolo fault (Fig. 9).
Based on the Late Permian horizon, the synform is defined by a 5.7km-amplitude parabolic flexure
across 170 km of wavelength, yielding an effective shortening of less than 0.30%.
Therefore, the strike-slip movement of the Nunkolo fault seems to have had more of a significant
transtensional component, rather than transpressional during the opening of this sub-basin at this
time under a wrench-dominated transtensional deformation.
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
The main consequence of the opening during the Middle-Late Triassic under this transtensional
deformation is the thickness variability of the Triassic sediments (Fig. 9). The Middle Triassic depicts
a slight thickening, especially at the top of the horizon. The thickening of the Late Triassic package is
more remarkable in the depocentre of this sub-basin (Fig. 9).
The Early–Middle Jurassic package follows the same pattern as described for the Late Triassic
horizon, although the progressive thickening is slightly less remarkable, and there is no thinning into
the WSW margin of this sub-basin, as shown in the Triassic strata. This could be the consequence of
the slow-down stage and infilling of the sub-basin during the Middle Jurassic (Fig. 9).
PT
In general terms, the isopach map of Jurassic sediments shapes the sub-basin and its SW-trending
driven by the Nunkolo transtensional strike-slip fault. With a relatively thin section, the Late Jurassic
appears to have been strongly affected by erosion, or not deposited, depicting the break-up
RI
unconformity or known as well as the Latest Callovian unconformity (Fig. 10). On the other hand, the
Early-Middle Jurassic shows a more complete record (Fig. 9).
SC
The Lower Cretaceous package, which mainly comprises the Echuca Shoals Formation, shows a
random thickness distribution with 48m encountered in Belalang-1. However, these Lower
U
Cretaceous sediments can be seen topping the Latest Callovian unconformity or filling half-grabens
to level a previous irregular palaeo-topography (Figs 9, 10).
N
A
The western footwall is made up of several horsts and grabens and is much more faulted than the
western West Timor sub-basin. These normal Mesozoic faults may have throws of more than one
M
The Jurassic appears faulted by normal faults and sometimes eroded (Figs 9, 10). Preservation of the
ED
Jurassic record seems restricted to Early–Middle Jurassic half-graben pockets, creating a scattered
Late Jurassic thickness distribution pattern (Figs 9, 10). The seismic lines nicely show an angular
unconformity with variable erosional rates. Even in some areas, such as Napoleon-1, the Echuca
T
Shoals Formation overlies the Triassic sediments of the Challis Formation with the entire Jurassic
EP
section absent (Figs. 5, 9, 10). Moreover, the Late Paleocene is not recorded in some horst blocks of
the western footwall and seems to have been either eroded or not deposited (Figs 9, 10). Here, the
angular unconformity is overlaid by a veneer of Eocene sediments.
CC
The Cova Lima sub-basin is governed by the NE-trending Manatuto transtensional strike-slip fault,
active since the Early–Middle Jurassic (Fig. 9). The sub-basin is a regional synform shaped by a 3.17
km amplitude parabolic flexure across 104 km of wavelength, yielding an effective shortening of
0.25%, very similar to the West Timor Sub-basin.
Therefore, the strike-slip movement of the Manatuto fault seems to have had much more of a
significant transtensional component, rather than transpressional during the opening of this sub-
basin (Fig. 9). Then, as well as the Manatuto fault, Nunkolo fault movement is defined under a
wrench-dominated transtensional deformation.
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
However, the transtensional deformation undergone during the Early–Middle Jurassic by the
Manatuto fault could be larger towards the northeast, but the Cova Lima Sub-basin is partially
covered by the orogenic wedge and makes difficult to calculate the extensional component in other
areas of the survey.
Similar to the West Timor Sub-basin, Cova Lima shows a double tectonic vergence WSW-ENE, being
the middle point at the depocenter of this sub-basin (Fig. 9). Neither Permian nor Triassic sediments
thicken in the Cova Lima sub-basin. Therefore, the transtensional sinistral strike-slip fault acted later
than the Late Triassic (Fig. 9). In both flanks of this sub-basin a positive flower structure can be
described, limiting this sub-basin (Fig. 9).
PT
The noticeable thickening for the Early–Middle Jurassic package at the depocenter has been
interpreted as the onset of the infill of this sub-basin, driven by the Manatuto fault. However, this
RI
package thins into the two antiforms, which confine the sub-basin (Fig. 9). The isopach map of the
Jurassic sediments outlines the SW-trending shape of this sub-basin driven by the Manatuto strike-
SC
slip fault. With a relatively thin layer, the Late Jurassic appears to have been strongly affected by
erosion or probably as well, no deposition characterizing the break-up unconformity, whereas the
Early–Middle Jurassic shows a more complete record (Figs 9, 10).
U
The Manatuto footwall corresponds to a palaeogeographic high with a much smaller fault
N
population, barely affecting the Permian sediments. The palaeo-high, fractured by Mesozoic normal
A
faults, seems to have been present since the middle Permian (Fig. 9).
M
Unlike the Permian, the Early–Middle Triassic package shows a clear thinning toward this paleo-high,
from 1680m nearby the fault plane to less than 300m recorded in the northeast margin of the
survey. This thinning is even more noticeable for the Late Triassic package, which barely reaches a
ED
In addition, the Jurassic appears to have been strongly affected by the presence of this
T
palaeogeographic high. Basically, the Jurassic record completely covers this paleo-high but shows a
EP
clear thinning to the NE, smoothing the relic palaeotopography. The Early–Middle Jurassic package is
thinner than the Late Jurassic, likely due to the post latest Callovian sea-level rise, which probably
created more accommodation space for sedimentation (Fig. 9).
CC
The Early Cretaceous, which comprises the Echuca Shoals Formation and its lateral equivalent
Darwin Formation, does not show a substantial thickness and seems to be confined in small grabens
A
showing a random thickness distribution. However, there is a slight increase in the Cova Lima along
the sub-basin axis, and a clear thinning above the palaeogeographic high (Fig. 9). In this paper, this
feature has been termed as the Timor Plateau because of its strong influence since the Middle-Late
Permian and recent tectonic activity reflected by the present-day bathymetric expression (Fig. 9).
Benincasa et al. (2012) suggested from onshore stratigraphic evidences a possible continental
plateau resembling others in the Northwest Shelf as a likely obstacle for the ongoing subduction.
Most probably, this Timor Plateau is the seismic evidence of this obstacle.
Structural Framework
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
As described previously, Timor Island can be defined as a classical arc-continent collision with
subsequent accretion (Fig. 1). The onset of the orogenic wedge was triggered because subduction
was primary jammed as thicker continental crust arrived at the subduction zone.
The thickening of the transitional crust progressively was locking subduction, given rise to the
present-day collision and subsequent fold-and-thrust belt. In more detail, the oblique collision
imprinted onshore a distinctive structural signature as described for several authors.
Many authors (Benincasa et al., 2012; Duffy et al., 2013) describe East Timor as dominated by recent
strike-slip faults that have been active into the Plio-Pleistocene and control the present-day
PT
topography of the Timor Island. The deformation style is described as a lateral crustal extrusion sub-
parallel to the Banda Arc with major orogen-parallel ENE-oriented faults on the northern and
southern sides of Timor displaying normal-sinistral and normal-dextral kinematics, respectively.
RI
Onshore, many topographic features show the distinctive characteristics of pull-apart basins such as
SC
the rhomboidal geometry of Lake Iralalaru or the pop-up structure of Mount Mundo Perdido. The
Mount Mundo Perdido area could have been developed at restraining bends in an E-W oriented
zone of sinistral strike-slip (Benincasa et al., 2012) and be one of the large massifs that throughout
U
the island are bound by NW-SE oriented dextral-normal faults and NE-SW oriented sinistral-normal
faults described by Duffy et al. (2013). The focal mechanisms in the area show a high population of
N
strike-slip solution.
A
This scenario could be labelled as a forearc sliver which already was introduced by Porritt et al.
M
Hence, the structural framework described onshore suggests that the transtensional convergent
ED
margin deformation is one of the main consequences of the geodynamics within the collisional zone
between the Australian and Eurasian/Pacific plates in the Timor region (Benincasa et al., 2012).
T
At present, the island continues to undergo uplift of around 1.5 mm/year (Audley-Charles, 1986).
EP
seismic interpretation
A
In the study area, the main tectonic event is the Neogene convergence of the Australia Plate and
Southeast Asian microplates. This collision induced the development of the Timor Trough and the
associated downwarping of the Australian Plate due to the vertical loading of the orogenic wedge of
Timor (Fig. 4).
Both, collision and flexural downwarping of the Australian plate involved widespread fault
reactivation across the western Bonaparte Basin. According to FrOG Tech Pty Ltd. (2005), this fault
reactivation occurred during different pulses of deformation in the Early Miocene, Late Miocene and
Late–Early Pliocene. Seismic images show present-day reactivation of Jurassic structures within the
flexure zone on the underplated Australian crust within an overall sinistral transtensional convergent
margin.
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
The strike-slip signature of the Nunkolo and Manatuto faults is seismically exposed as a double
tectonic vergence of the normal faults within the sedimentary package of the sub-basins, having a
change of polarity at their depocenter (Fig. 9).
The location of these features along the Australian continental edge plays a dominant role in
controlling the structural style and shaping of the Timor orogenic wedge. The Nunkolo footwall and
PT
the Timor Plateau represent a comparatively much larger obstacle to the advance of the thrust front
than the sub-basins. In the outer ridge, opposite to the West Timor and Cova Lima sub-basins,
collisional structures are less prominent, and the advance of the orogenic building is less restricted
RI
(Fig. 9).
SC
As mentioned previously, seismic interpretation shows that the Timor Plateau has been an
important palaeohigh since the Middle Permian and the Nunkolo footwall since the Middle to Late
Triassic (Fig. 9).
U
The Nunkolo footwall is less prominent than the Timor Plateau. It could have acted as a hinge joint
N
during the coeval sinistral movement between Nunkolo and Manatuto from Late Triassic to Middle
A
Jurassic (Fig. 9). The Marlborough fault system, South Island of New Zealand, is an analogue. This
fault system reflected southward migration of transform motion acting as an active Neogene
M
mechanical wedge with uplifting and rotation of their different domains (Little & Jones, 1998;
Reyners, 2013; Mortimer, 2014).
ED
The vertical loading and advance of the incoming orogenic wedge are the main driving forces for
building the orogenic wedge (Figs. 4, 6, 7, 11). It seems that the gravitational-driven process is
secondary force acting in the construction of the orogenic building but more thorough studies of the
T
dynamics of this orogenic wedge are needed to quantify both processes and corroborate this
EP
assessment.
Two styles of post-Miocene deformation are present: shallow thin-skinned and deep-seated
CC
deformation (Figs. 6, 7, 8). The shallow thin-skinned deformation creates a well-defined shallow fold-
and-imbricated-thrust belt developed in the passive margin successions of the Australian Plate
(Gondawana Megasequence) termed here “shallow thin-skinned tectonics” (Fig. 6). The inverted,
A
deeper, inherited Late Carboniferous-Permian structures form thick fold-and-thrust sheets detached
on the Late Permian and Early Triassic shales, termed “deep tectonics” (Fig. 8).
The shallow, thin-skinned deformation is developed by two detachments corresponding to the Top
Miocene and Late Cretaceous shales, whereas the deep-seated tectonics is focused on the inversion
of the inherited Mesozoic–Palaeozoic horsts and grabens, some of which are probably basement-
involved faults (Figs. 6, 7, 8).
The mechanical properties of the Australian continental crust preconditioned the development of
the thin-skinned tectonics along the Timor orogenic wedge. Two levels of shallow detachments can
be recognized across the study area: Late Cretaceous shales overlying the Late Callovian break-up
unconformity and the Top Miocene reflectors.
On the other hand, it is worth noting that the mechanical properties or thickness variations of the
trench sediments along-strike do not seem to have predisposed the shallow deformation or worked
as detachment levels (Fig. 9).
The shallow thin-skinned deformation is determined by two drivers: the compressional horizontal
PT
stress provoked by the southward overthrusting of the outer wedge and the gravitational
component caused by internal stress release to regain or to maintain stability within the outer
wedge (Figs. 6, 7, 8, 11).
RI
The latter occurs in the first few kilometers upslope of tectonically active areas of the outer wedge.
SC
Thrusts develop at the toes of the sliding sheet affecting Neogene deposits. The southerly frontal
accretion of the orogenic wedge incorporates trench sediments into the thrust belt. In some dip-
lines opposite to the sub-basins, the outer wedge does not display this gravitational-driven process
U
and shows undeformed Plio-Pleistocene sediments veneering the slope of the wedge. In contrast,
this gravitational signature is recorded in the dip-lines opposite to the Nunkolo footwall and Timor
N
Plateau (Figs. 6, 7, 9, 11).
A
The Late Cenozoic succession in the Australian margin shows a variable thickness depending on the
M
activity of the main NW-SE strike-slip faults (Fig. 9). This variability has a strong influence on the
position of the detachment in the orogenic wedge and its style of deformation. A thin Late Cenozoic
succession allows the development of a low-angle detachment on which imbricated thrusting is the
ED
most common structural style. This low-angle detachment can be extended as far as 20 km
northward in the front-of-deformation with no influences of the deeper Mesozoic nappes. This thin-
T
skinned tectonic style with detachment on the Top Miocene reflector has a similar structural style to
the thin-skinned tectonic developed on the Callovian seismic reflector (Figs. 4, 6, 7).
EP
On the other hand, the increase of Neogene thickness tends to obstruct the downwards propagation
of the wedge deformation and build a complex and deeper accretionary wedge (Fig. 9). Hence, when
CC
the Late Cenozoic–Quaternary section is thicker than two kilometers, it makes an irregular and
stepped frontal ramp, detached on the Callovian unconformity (Figs. 6, 7). This irregular propagation
is also visible on the inverted grabens and on the top of deeper fault-propagation folds. When the
A
Callovian unconformity reaches 5 km depth or more, imbricated structures are developed on the
fore-thrust sheet and the thickness of the orogenic wedge increases, developing a new detachment
on the top of the Miocene (Fig. 9).
The Late Callovian unconformity reflector is the most important detachment in the study area when
the driver of the deformation is the compressional horizontal stress due to the southward advance
of the wedge (Fig. 10). Opposite the Nunkolo footwall, the main detachment becomes the top
Miocene reflector (Fig. 8).
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
This tectonic unit below the Callovian unconformity is a complex succession of imbricated foreland-
verging thrust sheets of Mesozoic and Cenozoic units. Thrust faults nucleate in a forward-breaking or
piggy-back sequence in the southward direction of thrust transport, with younger thrusts developing
in front of older thrusts (Figs. 4, 6, 7).
Deeper inverted grabens induce disruption in the thin-skinned tectonic by breaking up the
detachment surface. This disruption generates a stepwise detachment with local backward tilting
where imbricated thrusts and antiformal stacks are developed.
To the west margin of Timor, deformation becomes more complex due to the inversion of deep
PT
Jurassic and Paleozoic structures (Fig. 8). Also, the fold-and-thrust belt is getting closer to the suture
zone which is running NE-SW to the north of Leti Island. At this point, deep metamorphic nappes are
already exhumed, merging the post-Miocene sequences in a strong imbrication of thrust sheets and
RI
coeval folding. Asymmetric synclines are overthrust and tightening, finishing in prominent back-
thrusts.
SC
The synchronous inversion of the inherited deep-seated fractures during the evolution of the thin-
skinned deformation is evidenced by the presence of decapitated normal faults forming the deeper
structures.
U
N
Deep tectonics
A
Normal reactivation and over-thrusting of inherited Mesozoic and Palaeozoic structures from the
M
Gondwana Megasequence result in a prominent deformation in the frontal part of the orogenic
wedge. This deep-seated deformation disturbs shallow detachments by locking and decoupling.
ED
The main consequence is the generation of thrust sheets developed on the Triassic and Early–Middle
Jurassic successions (Fig. 8). The broadband seismic data shows three main structural styles of deep
structures: reactivation of deep-seated graben structures, short-cutting faults and deep-fold
T
propagational faults.
EP
Reactivation of deep-seated graben structures generates the most common geometry along the
outer wedge. Originally, these reactivated structures were developed during the NW-SE
CC
“Westralian” extensional event, as the result of the reactivation of Proterozoic and older Palaeozoic
NE-trending faults. Later, these Proterozoic and Palaeozoic faults were developed as normal faults in
the Late Callovian during the break-up of Gondwana and finally reactivated during the Late Miocene
A
collision. This linkage, tectonic inheritance and overprinting played a paramount role in the study
area and most probably throughout the Northwest Shelf of Australia (Lister et al., 1991; Harrowfield
et al., 2005). Ambient noise tomography studies in the Timor Trough have recognised underlying
structural features in the lowermost crust and uppermost mantle as main drivers of the upper
crustal shortening (Porritt et al., 2016).
Depending on the height of incoming structures to the trough, in relation to the southward
horizontal motion component of the wedge, Mesozoic faults can be inverted or normally reactivated
(Figs. 6, 7, 11). The height of these structures nearby the trough depends on the amount of
downward flexing before reaching the trench at that point. Moreover, the dipping orientation of
incoming normal faults into the trench plays an important role.
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
Another consequence of the downward flexure of the Australian plate is the NNW extensional stress
provoking the generation of Neogene normal faults affecting Miocene and younger packages.
Sometimes, these Neogene faults are linked with the Mesozoic faults (Fig. 4). The relationship
between the strike of the Jurassic faults and the horizontal maximum stress (σHmax) plays an
important role in how the reactivated Jurassic faults can be linked with the newly developed
Neogene faults. Low angles (<20º) between Jurassic fault strikes and σHmax generate hard-linkages,
whereas soft-linkages are likely to occur above that threshold (Langhi et al., 2011).
The inversion of the SE-dipping normal faults develops complex structures in the front of the outer
PT
wedge, typically decoupling and locking shallower thrusts. Measurements of critical taper angles (
and ) in the seismic dip-lines show a clear increase in these areas, especially in the Timor Plateau.
The occurrence of this deep-seated deformation is higher in areas opposite to the Nunkolo footwall
RI
and Timor Plateau (Figs. 6, 7, 8, 9).
SC
Reactivation of NW-dipping normal faults is observed on the easternmost part of the survey. This
reactivation results in deep-seated high-angle reverse faults and the development of inversion
anticlines in the Late Mesozoic and Early Cainozoic (Fig. 8). The inversion is mainly along the master
U
fault and synthetic faults of the Jurassic grabens. This scenario develops important and complex
deformation on the thin-skinned tectonics to the inner part of the orogenic wedge.
N
A
Opposite the sub-basins, pre-existing normal faults show a deep, low-angle detachment on the top
of the Permian reflector. Therefore, inversion results in a mild deformation of the overlying
M
Finally, horst inversion also occurs in the downward flexure of the Australian Plate opposite to the
ED
highest area of the Timor Plateau, beside the Manatuto fault surface. In this area, both faults of
several horsts have been inverted creating a bumpy bathymetry along-strike of the inverted faults
(Fig. 11).
T
EP
Short-cutting reverse faults are developed in a more advanced stage of inversion on SE-dipping
faults. This process takes place at the front of the deformation and is linked with the development of
thin skinned detachments (Figs 6, 7). This mechanism of deformation has been described in the
CC
Central Apennines where Mesozoic–Cainozoic normal faults are decapitated thrusts (Scisciani, 2009;
Bonini et al., 2012).
A
This style is mostly described by the presence of the Nunkolo footwall and the Timor Plateau, which
provokes the inversion of the SE-dipping normal faults, developing complex structures in front of the
outer wedge and typically decoupling and locking shallow decollements (Fig. 9). Measurements of
critical taper angles ( and ) in the seismic dip-lines show a clear increase in these areas, especially
in the Timor Plateau.
The seismic data images several examples of SE-dipping faults that are bent due to the advance,
weight and friction of the over-thrusting orogenic wedge. This results in arcuate-shaped inverted
faults and horsts that have been “decapitated or truncated” by low angle NNW-dipping thrusts
detaching on Early Triassic shale. The length and strike orientation of these decapitated faults seem
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
to control the length and orientation of the ramps of these thrusts. Hence, the scraped-off Triassic
and Jurassic potential reservoirs and source rock levels are constantly uplifted by thrusting and
stacked. These structures are more developed opposite to the palaeographic features, increasing the
topographic slope of the wedge (Figs. 6-7-8).
Sometimes, this situation can also be described in the NW-dipping faults, if the downward flexure of
the Australian plate does not place these horsts below the horizontal compressional stress resulting
from the advance of the orogenic wedge (Fig. 11). Most commonly, the NW-dipping faults are
reactivated as normal faults due to the vertical loading of the incoming wedge bringing down the
graben to within the hydrocarbon window, forming interesting kitchen areas for petroleum
PT
exploration.
Deep fault-propagation folds are observed mainly in the Triassic shales, with a possible detachment
RI
on the top of the Late Permian reflector. These folds are mainly south-verging with a lower angle
than the inverted normal faults (Figs. 8).
SC
In some dip-lines, these structures are related with the most recent deformation which in some
cases is located right below the current foredeep.
U
There are no signs of deep fault propagation folds nor inversion of inherited normal faults
N
southward the foredeep in the Australian Platform. It may imply that the thrusts detached at the top
A
of the Late Permian are currently inactive.
M
Geographical analogues
ED
The Timor accretionary wedge is probably one of the best examples in world of an early-stage
T
oblique arc-continent collision (Keep & Haig, 2010). Another world-class examples such as Taiwan,
EP
Papua New Guinea and the Urals are in a more advanced stages giving a better understanding about
the nature and evolution of their structural styles and kinematics of their angular collision. In
particular, the Taiwan orogeny has a similar geodynamic evolution and structural configuration as
CC
Timor Orogeny (Huang et al., 2000) showing a fast highly oblique collision NW-SE at ~70 mm/yr
exhuming diachronically the accretionary wedge in a complex fold-and-thrust belt (Chi et al., 1981).
Analogues experiments of the Taiwan orogeny show that the rotational and compressional
A
kinematics result in a local partitioning between thrusting and wrenching, where faults or shear
zones can be rotated around the indentation point of the backstop by transcurrent and bookshelf
faulting (Chia-Yu Lu & Malavieille, 1994). Additionally, it is suggested that the influence of inherited
relieve and nature of the passive margin may play an important role on the kinematic evolution of
young arc-continent collisional orogens (Malavieille & Trullenque, 2009) as described in the present
work for Timor Orogeny.
Importance of the inherited passive margin structures involved in young fold-and-thrust belts.
Mention to Central Apennines, Alps and Central Andes
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
The importance of the inversion of the inherited normal faults and its kinematics have been
particularly well documented in the several inverted basins around the world like the Alps, Central
Apennines and Central Andes (Ghisetti & Vezzani, 2002; Butler et al., 2006; Bonini et al., 2012). In
contractional contexts, as Timor Orogeny, the inherited configuration of the passive margins plays a
key role on the nature and evolution of the structural styles. For instance, the development of
multiple detachment zones, rotational grabens, shortcut thrusting and the ultimate decapitation of
normal faults, have been reported as typical structural styles of inverted passive margins in
collisional settings in the Central Apennines and the Atacama Region (Scisciani, 2009; Bonini et al.,
2012). These typical structures were also developed during the inversion of the Timor accretionary
wedge indicating the importance of the passive structural inheritance on the definition of the
PT
present-day structural styles.
Conclusions
RI
SC
The interpretation and analysis of the broadband BandaSeis 2D seismic data brings new insights
about the geology of the colliding zone between the downwardly Australian continental margin and
the over-thrusting orogenic wedge of Timor.
U
Moreover, the identification of the main mechanisms governing the outer ridge of Timor gives clues
N
to understand the whole process of orogenic building.
A
Several geological events such as Pillara, Westralian, Fitzroy and finally the Latest Callovian break-up
M
have imprinted a very distinctive signature to this inherited paleogeography with key features to
consider.
ED
The main geological structures governed by the Nunkolo and Manatuto faults are the West Timor
and Cova Lima sub-basins, the Timor Plateau and the Nunkolo footwall (Fig. 9).
T
These geological features play an important role controlling the style of collision between Australia
EP
and Sundaland and shaping the internal architecture of the outer orogenic wedge of Timor (Fig. 9).
The Timor Plateau is the most striking feature when looking at the 750 km W-E cross-section along-
CC
strike of the Timor Trough (Fig. 9). Working as a structural indenter, the Timor Plateau imprints more
deep-seated tectonics on the outer ridge in the east of Timor Island. The structural style changes
westward, moving away from the Timor Plateau into the Cova Lima sub-basin. Both sub-basins make
A
overthrusting easier for the outer wedge, comparable to paving a path. Therefore, the shallow
tectonic signature changes into the shallow thin-skinned tectonics (Fig. 9).
The Nunkolo footwall is less overwhelming obstacle and does not impede subduction as much as the
Timor Plateau. However, it has enough entity to switch the internal architecture of the outer wedge
to deep-seated tectonics from the shallow thin-skinned tectonics described in the area of influence
of the West Timor and Cova Lima sub-basins (Fig. 9).
Seismic evidences underpin the Timor Plateau as a tectonic indenter and allow us to discard the slab
detachment hypothesis. The Timor Plateau or Nunkolo footwall fit much better with the short
wavelength pattern described in several surveys, diachronism and differential uplift rates. However,
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
further studies based on critical taper analysis are needed to understand the outer wedge stability
and the role played by the Timor Plateau as an indenter (Fig. 9).
The basement-involved Nunkolo and Manatuto transtensional strike-slip faults not only helped
shaped the Australian continental margin but also compartmentalised the sedimentary input into
the study area during Triassic and Jurassic times playing an important role in facies distribution.
Acknowledgements
We thank the governments of Indonesia and Timor-Leste for their continued support and CGG for
PT
permission to publish this paper. Thanks to Jo Firth for comments on the manuscript and to Robert
Hall (Royal Holloway, University of London) for fruitful discussions about the structural evolution of
RI
the Banda region.
SC
References
U
Audley-Charles, M.G. 1986. Rates of Neogene and Quaternary tectonic movements in the Southern
Banda Arc based on micropalaeontology. Journal of the Geological Society, 143, 161–175,
N
https://doi.org/10.1144/gsjgs.143.1.0161.
A
Audley-Charles, M.G. 2004. Ocean trench blocked and obliterated by Banda forearc collision with
M
Audley-Charles, M.G. 2011. Tectonic post-collision processes in Timor. Geological Society of London,
Special publications, 355, 241–266, https://doi.org/10.1144/SP355.12.
T
Audley-Charles, M.G., Ballantyne, P.D. & Hall, R. 1988. Mesozoic-Cenozoic rift-drift sequence of
Asian fragments from Gondwanaland. Tectonophysics, 155, 317–330,
EP
https://doi.org/10.1016/0040-1951(88)90272-7.
CC
Belgarde, C., Manatschal, G., Kusznir, N., Scarselli, S. & Ruder, M. 2015. Rift processes in the
Westralian Superbasin, North West Shelf, Australia: insights from 2D deep reflection seismic
interpretation and potential fields modelling. The APPEA Journal, 55, 400,
A
https://doi.org/10.1071/AJ14035.
Benincasa, A., Keep, M. & Haig, D.W. 2012. A restraining bend in a young collisional margin: Mount
Mundo Perdido, East Timor. Australian Journal of Earth Sciences, 59, 859–876,
https://doi.org/10.1080/08120099.2012.686453.
Bird, P. 2003. An updated digital model of plate boundaries. Geochemistry, Geophysics, Geosystems,
4, https://doi.org/10.1029/2001GC000252.
Bock, Y., Prawirodirdjo, L., et al. 2003. Crustal motion in Indonesia from Global Positioning System
measurements. Journal of Geophysical Research, 108, 2367,
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
https://doi.org/10.1029/2001jb000324.
Bonini, M., Sani, F. & Antonielli, B. 2012. Basin inversion and contractional reactivation of inherited
normal faults: A review based on previous and new experimental models. Tectonophysics, 522–
523, 55–88, https://doi.org/10.1016/j.tecto.2011.11.014.
Butler, R.W.H., Tavarnelli, E. & Grasso, M. 2006. Structural inheritance in mountain belts: An Alpine-
Apennine perspective. Journal of Structural Geology, 28, 1893–1908,
https://doi.org/10.1016/j.jsg.2006.09.006.
Chamalaun, F.H. & Grady, A.E. 1978. The tectonic development of Timor: a new model and its
PT
implications for petroleum exploration. The APPEA Journal, 18, 102,
https://doi.org/10.1071/AJ77012.
RI
Charlton, T.R. 2002. The Petroleum Potential of East Timor. APPEA Journal, 351–369.
SC
Chi, W.R., Namson, J. & Suppe, J. 1981. Stratigraphic record of plate interactions in Coastal Range,
eastern Taiwan. Memoir of the Geological Society of China, 4, 155–194.
U
Chia-Yu Lu & Malavieille, J. 1994. Oblique convergence, indentation and rotation tectonics in the
N
Taiwan Mountain Belt: Insights from experimental modelling. Earth and Planetary Science
Letters, 121, 477–494, https://doi.org/10.1016/0012-821X(94)90085-X.
A
DeMets, C., Gordon, R.G., Argus, D.F. & Stein, S. 1990. Current plate motions. Geophysical Journal
M
DeMets, C., Gordon, R.G., Argus, D.F. & Stein, S. 1994. Effect of recent revisions to the geomagnetic
reversal time scale on estimates of current plate motions. Geophysical Research Letters, 21,
2191–2194, https://doi.org/10.1029/94GL02118.
T
DeMets, C., Gordon, R.G. & Argus, D.F. 2010. Geologically current plate motions. Geophysical Journal
EP
Duffy, B., Quigley, M.C., Harris, R.A. & Ring, U. 2013. Arc-parallel extrusion of the Timor sector of the
CC
Ely, K.S., Sandiford, M., Hawke, M.L., Phillips, D., Quigley, M.C. & dos Reis, J.E. 2011. Evolution of
A
Ataúro Island: Temporal constraints on subduction processes beneath the Wetar zone, Banda
Arc. Journal of Asian Earth Sciences, 41, 477–493,
https://doi.org/10.1016/j.jseaes.2011.01.019.
England, P., Engdahl, R. & Thatcher, W. 2004. Systematic variation in the depths of slabs beneath arc
volcanoes. Geophysical Journal International, 156, 377–408, https://doi.org/10.1111/j.1365-
246X.2003.02132.x.
Eva, C., Cattaneo, M. & Merlanti, F. 1988. Seismotectonics of the central segment of the Indonesian
Arc. Tectonophysics, 146, 241–259, https://doi.org/10.1016/0040-1951(88)90094-7.
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
FrOG Tech Pty Ltd. 2005. OZ SEEBASETM Study 2005, Public Domain Report to Shell Development
Australia.
Genrich, J.F., Bock, Y., McCaffrey, R., Calais, E., Stevens, C.W. & Subarya, C. 1996. Accretion of the
southern Banda arc to the Australian plate margin determined by Global Positioning System
measurements. Tectonics, 15, 288–295, https://doi.org/10.1029/95TC03850.
Ghisetti, F. & Vezzani, L. 2002. Normal faulting, extension and uplift in the outer thrust belt of the
central Apennines (Italy): role of the Caramanico fault. Basin Research, 14, 225–236,
https://doi.org/10.1046/j.1365-2117.2002.00171.x.
PT
Gill, J.B. 1981. Orogenic Andesites and Plate Tectonics. Orogenic Andesites and Plate Tectonics, 16,
390, https://doi.org/10.1007/978-3-642-68012-0.
RI
Haig, D.W. 2012. Palaeobathymetric gradients across Timor during 5.7–3.3 Ma (latest Miocene–
Pliocene) and implications for collision uplift. Palaeogeography, Palaeoclimatology,
SC
Palaeoecology, 331–332, 50–59, https://doi.org/10.1016/j.palaeo.2012.02.032.
U
Haig, D.W. & McCartain, E.W. 2007. Carbonate pelagites in the post-Gondwana succession
(Cretaceous – Neogene) of East Timor. Australian Journal of Earth Sciences, 54, 875–897,
N
https://doi.org/10.1080/08120090701392739.
A
Haig, D.W. & McCartain, E.W. 2010. Triassic Organic-Cemented Siliceous Agglutinated Foraminifera
M
Haig, D.W., McCartain, E.W., Barber, L. & Backhouse, J. 2007. Triassic-Lower Jurassic Foraminiferal
Indices for Bahaman-Type Carbonate-Bank Limestones, Cablac Mountain, East Timor. The
Journal of Foraminiferal Research, 37, 248–264, https://doi.org/10.2113/gsjfr.37.3.248.
T
Haig, D.W., McCartain, E.W., Keep, M. & Barber, L. 2008. Re-evaluation of the Cablac Limestone at its
EP
type area, East Timor: Revision of the Miocene stratigraphy of Timor. Journal of Asian Earth
Sciences, 33, 366–378, https://doi.org/10.1016/j.jseaes.2008.03.002.
CC
Hall, R. 2002. Cenozoic geological and plate tectonic evolution of SE Asia and the SW Pacific:
Computer-based reconstructions, model and animations. Journal of Asian Earth Sciences, 20,
A
353–431, https://doi.org/10.1016/S1367-9120(01)00069-4.
Hall, R. 2011. Australia–SE Asia collision: plate tectonics and crustal flow. Geological Society of
London, Special publications, 355, 75–109, https://doi.org/10.1144/SP355.5.
Hall, R. 2012. Late Jurassic-Cenozoic reconstructions of the Indonesian region and the Indian Ocean.
Tectonophysics, 570–571, 1–41, https://doi.org/10.1016/j.tecto.2012.04.021.
Hall, R., Cottam, M.A. & Wilson, M.E.J. 2011. The SE Asian Gateway: History and Tectonics of the
Australia-Asia Collision, https://doi.org/10.1144/SP355.1.
Hamilton, W. 1979. Tectonics of the Indonesian Region. Geological Survey Professional Paper, 1078,
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
308, https://doi.org/10.1016/0003-6870(73)90259-7.
Haq, B.U., Hardenbol, J., Vail, P.R., Haq, B.U., Hardenbol, J.A.N. & Vail, P.R. 1987. Chronology of
Fluctuating Sea Levels Since the Triassic. Science, 235, 1156–1167.
Harris, R.A. 2006. Rise and fall of the Eastern Great Indonesian arc recorded by the assembly,
dispersion and accretion of the Banda Terrane, Timor. Gondwana Research, 10, 207–231,
https://doi.org/10.1016/j.gr.2006.05.010.
Harris, R.A., Sawyer, R.K. & Audley-Charles, M.G. 1998. Collisional melange development: geologic
associations of active melange-forming processes with exhumed melange facies in the western
PT
Banda orogen, Indonesia. Tectonics, 17, 458–479, https://doi.org/10.1029/97TC03083.
RI
Harrowfield, M., Holdgate, G.R., Wilson, C.J.L. & Mcloughlin, S. 2005. Tectonic significance of the
Lambert Graben, East Antarctica: reconstructing the Gondwanan rift. Geology, 33, 197–200.
SC
Hartono, H.M.S. 1990. Late Cenozoic tectonic development of the Southeast Asian continental
margin in the Banda Sea area. Tectonophysics, 181, 267–276, https://doi.org/10.1016/0040-
1951(90)90020-9.
U
N
Heine, C. & Müller, R.D. 2005. Late Jurassic rifting along the Australian North West Shelf: margin
geometry and spreading ridge configuration. Australian Journal of Earth Sciences, 52, 27–39,
A
https://doi.org/10.1080/08120090500100077.
M
Heine, C., Quevedo, L., McKay, H. & Müller, R.D. 2012. Plate Tectonic Consequences of competing
models for the origin and history of the Banda Sea subducted oceanic lithosphere. Eastern
ED
Herrington, R.J., Scotney, P.M., Roberts, S., Boyce, A.J. & Harrison, D. 2011. Temporal association of
T
593, https://doi.org/10.1016/j.gr.2010.10.011.
CC
Hilton, D.R. & Craig, H. 1989. A helium isotope transect along the Indonesian archipelago. Nature,
342, 906–908, https://doi.org/10.1038/342906a0.
Hilton, D.R., Hoogewerff, J.A., van Bergen, M.J. & Hammerschmidt, K. 1992. Mapping magma
A
sources in the east Sunda-Banda arcs, Indonesia: Constraints from helium isotopes. Geochimica
et Cosmochimica Acta, 56, 851–859, https://doi.org/10.1016/0016-7037(92)90105-R.
Hinschberger, F., Malod, J.-A., Réhault, J.-P., Villeneuve, M., Royer, J.-Y. & Burhanuddin, S. 2005. Late
Cenozoic geodynamic evolution of eastern Indonesia. Tectonophysics, 404, 91–118,
https://doi.org/10.1016/j.tecto.2005.05.005.
Huang, C.Y., Yuan, P.B., Lin, C.W., Wang, T.K. & Chang, C.P. 2000. Geodynamic processes of Taiwan
arc-continent collision and comparison with analogs in Timor, Papua New Guinea, Urals and
Corsica. Tectonophysics, 325, 1–21, https://doi.org/10.1016/S0040-1951(00)00128-1.
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
Keep, M. & Haig, D.W. 2010. Deformation and exhumation in Timor: Distinct stages of a young
orogeny. Tectonophysics, 483, 93–111, https://doi.org/10.1016/j.tecto.2009.11.018.
Keep, M., Longley, I.M. & Jones, R. 2003. Sumba and its effect on Australia’s northwestern margin.
In: Evolution and Dynamics of the Australian Plate. Geological Society of America, 309–312.,
https://doi.org/10.1130/0-8137-2372-8.309.
Kelman, A.P., Kennard, J.M., et al. 2014. Bonaparte Basin Biozonation and Stratigraphy, Chart 33.
Koulali, A., McClusky, S., et al. 2017. The kinematics of crustal deformation in Java from GPS
observations: Implications for fault slip partitioning. Earth and Planetary Science Letters, 458,
PT
69–79, https://doi.org/10.1016/j.epsl.2016.10.039.
RI
Kreemer, C., Holt, W.E., Goes, S. & Govers, R. 2000. Active deformation in eastern Indonesia and the
Philippines from GPS and seismicity data. Journal of Geophysical Research: Solid Earth, 105,
663–680, https://doi.org/10.1029/1999JB900356.
SC
Kreemer, C., Blewitt, G. & Klein, E.C. 2014. A geodetic plate motion and Global Strain Rate Model.
Geochemistry, Geophysics, Geosystems, 15, 3849–3889,
https://doi.org/10.1002/2014GC005407. U
N
Langhi, L., Ciftci, N.B. & Borel, G.D. 2011. Impact of lithospheric flexure on the evolution of shallow
A
faults in the Timor foreland system. Marine Geology, 284, 40–54,
M
https://doi.org/10.1016/j.margeo.2011.03.007.
Lister, G.S., Etheridge, M.A. & Symonds, P.A. 1991. Detachment models for the formation of passive
ED
Little, T.A. & Jones, A. 1998. Seven million years of strike-slip and related off-fault deformation,
T
northeastern Marlborough fault system, South Island, New Zealand. Tectonics, 17, 285–302,
https://doi.org/10.1029/97TC03148.
EP
Longley, I.M., Buessenschuett, C., et al. 2002. The North West shelf of Australia - A woodside
CC
Malavieille, J. & Trullenque, G. 2009. Consequences of continental subduction on forearc basin and
A
McCaffrey, R. 1992. Oblique plate convergence, slip vectors, and forearc deformation. Journal of
Geophysical Research, 97, 8905–8915, https://doi.org/10.1029/92JB00483.
McCaffrey, R. 1996. Slip partitioning at convergent plate boundaries of SE Asia. Geological Society of
London, Special publications, 106, 3–18, https://doi.org/10.1144/GSL.SP.1996.106.01.02.
Metcalfe, I. 2009. Late Palaeozoic and Mesozoic tectonic and palaeogeographical evolution of SE
Asia. Geological Society of London, Special publications, 315, 7–23,
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
https://doi.org/10.1144/SP315.2.
Miller, K.G., Kominz, M.A., et al. 2005. The Phanerozoic Record of Global Sea-Level Change. Science,
310, 1293–1298, https://doi.org/10.1126/science.1116412.
Mortimer, N. 2014. The oroclinal bend in the South Island, New Zealand. Journal of Structural
Geology, 64, 32–38, https://doi.org/10.1016/j.jsg.2013.08.011.
Müller, C., Kopp, H., et al. 2008. From subduction to collision: The Sunda-Banda Arc transition. Eos,
89, 49–50, https://doi.org/10.1029/2008EO060001.
PT
Nebel, O., Vroon, P.Z., van Westrenen, W., Iizuka, T. & Davies, G.R. 2011. The effect of sediment
recycling in subduction zones on the Hf isotope character of new arc crust, Banda arc,
RI
Indonesia. Earth and Planetary Science Letters, 303, 240–250,
https://doi.org/10.1016/j.epsl.2010.12.053.
SC
Nugroho, H., Harris, R.A., Lestariya, A.W. & Maruf, B. 2009. Plate boundary reorganization in the
active Banda Arc-continent collision: Insights from new GPS measurements. Tectonophysics,
479, 52–65, https://doi.org/10.1016/j.tecto.2009.01.026.
U
N
Porritt, R.W., Miller, M.S., O’Driscoll, L.J., Harris, C.W., Roosmawati, N. & Teofilo da Costa, L. 2016.
Continent–arc collision in the Banda Arc imaged by ambient noise tomography. Earth and
A
Planetary Science Letters, 449, 246–258, https://doi.org/10.1016/j.epsl.2016.06.011.
M
Pryer, L., Blevin, J., et al. 2014. Structural architecture and basin evaluation of the North West Shelf.
The APPEA Journal, 54, 474, https://doi.org/10.1071/AJ13047.
ED
Reyners, M. 2013. The central role of the Hikurangi Plateau in the Cenozoic tectonics of New Zealand
and the Southwest Pacific. Earth and Planetary Science Letters, 361, 460–468,
T
https://doi.org/10.1016/j.epsl.2012.11.010.
EP
Richardson, A.N. & Blundell, D.J. 1996. Continental collision in the Banda arc. Geological Society of
London, Special publications, 106, 47–60, https://doi.org/10.1144/GSL.SP.1996.106.01.05.
CC
Roosmawati, N. & Harris, R.A. 2009. Surface uplift history of the incipient Banda arc-continent
collision: Geology and synorogenic foraminifera of Rote and Savu Islands, Indonesia.
Tectonophysics, 479, 95–110, https://doi.org/10.1016/j.tecto.2009.04.009.
A
Ryan, W.B.F., Carbotte, S.M., et al. 2009. Global Multi-Resolution Topography synthesis.
Geochemistry, Geophysics, Geosystems, 10, Q03014, https://doi.org/10.1029/2008GC002332.
Sani, K., Jacobson, M.I. & Sigit, R. 1995. The thin-skinned thrust structures of Timor. Proceedings of
the Indonesian Petroleum Association 24th Annual Convention, 277–293.
Scisciani, V. 2009. Styles of positive inversion tectonics in the Central Apennines and in the Adriatic
foreland: Implications for the evolution of the Apennine chain (Italy). Journal of Structural
Geology, 31, 1276–1294, https://doi.org/10.1016/j.jsg.2009.02.004.
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
Seton, M., Dietmar Müller, R., et al. 2012. Global continental and ocean basin reconstructions since
200 Ma. Earth-Science Reviews, 113, 212–270,
https://doi.org/10.1016/j.earscirev.2012.03.002.
Spakman, W. & Hall, R. 2010. Surface deformation and slab–mantle interaction during Banda arc
subduction rollback. Nature Geoscience, 3, 562–566, https://doi.org/10.1038/ngeo917.
Syracuse, E.M. & Abers, G.A. 2006. Global compilation of variations in slab depth beneath arc
PT
volcanoes and implications. Geochemistry, Geophysics, Geosystems, 7,
https://doi.org/10.1029/2005GC001045.
RI
Tate, G.W., McQuarrie, N., van Hinsbergen, D.J.J., Bakker, R.R., Harris, R.A. & Jiang, H. 2015. Australia
going down under: Quantifying continental subduction during arc-continent accretion in Timor-
SC
Leste. Geosphere, 11, 1860–1883, https://doi.org/10.1130/GES01144.1.
U
Vroon, P.Z., van Bergen, M.J., Klaver, G.J. & White, W.M. 1995. Strontium, neodymium, and lead
isotopic and trace-element signatures of the East indonesian sediments: provenance and
N
implications for banda arc magma genesis. Geochimica et Cosmochimica Acta, 59, 2573–2598,
https://doi.org/10.1016/0016-7037(95)00151-4.
A
M
Wheller, G.E., Varne, R., Foden, J.D. & Abbott, M.J. 1987. Geochemistry of quaternary volcanism in
the Sunda-Banda arc, Indonesia, and three-component genesis of island-arc basaltic magmas.
Journal of Volcanology and Geothermal Research, 32, 137–160, https://doi.org/10.1016/0377-
ED
0273(87)90041-2.
Wille, M., Nebel, O., Pettke, T., Vroon, P.Z., König, S. & Schoenberg, R. 2018. Molybdenum isotope
T
variations in calc-alkaline lavas from the Banda arc, Indonesia: Assessing the effect of crystal
EP
fractionation in creating isotopically heavy continental crust. Chemical Geology, 485, 1–13,
https://doi.org/10.1016/j.chemgeo.2018.02.037.
CC
Williams, S.E., Flament, N., Dietmar Müller, R. & Butterworth, N. 2015. Absolute plate motions since
130 Ma constrained by subduction zone kinematics. Earth and Planetary Science Letters, 418,
66–77, https://doi.org/10.1016/j.epsl.2015.02.026.
A
Zahirovic, S., Seton, M. & Müller, R.D. 2014. The Cretaceous and Cenozoic tectonic evolution of
Southeast Asia. Solid Earth, 5, 227–273, https://doi.org/10.5194/se-5-227-2014.
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
Figure captions
Fig. 1. Locality map with BandaSeis dataset indicated. Banda Arc regional map shows the main
regional geological structures such as the subduction of the Argo Abyssal Plain under Sunda Arc
following eastwardly the onset of the Banda Arc orogenic toe-thrust at south Sumba island. Labelled
in the map the shape and extent of the inner and outer arc as well as the foreland basin and the
volcanic arc. Underlying base map: Global Multi-Resolution Topography (GMRT) synthesis (Ryan et
al., 2009).
Fig. 2. Set of paleogeographic maps of the SE Asia and North West shelf since the Early Oligocene
PT
(modified from Hall 2002). These illustrations depict since 30 Ma the evolution of the subduction of
the Australian margin. Maps also outline the shape of the Australian margin before colliding with
Sundaland, considered to play an important role in the outcome of the collision.
RI
Fig. 3. Chronostratigraphic chart for the northern Bonaparte Basin based on the Bonaparte Basin
SC
Biozonation and Stratigraphy Chart 33 (Kelman et al., 2014). Sea-level curves based on Haq et al.
1987 and Miller et al. 2005 as well as short-term sea-level curve provided by Geoscience Australia.
U
Fig. 4. NNW-SSE 2D seismic line across Timor Trough with the geological interpretation (inset map
for location indicated by thicker pink line). Note thickening of the Early–Middle Jurassic package as
N
well as the constant thickness of the carbonate Late and Middle Permian succession (vertical scale in
A
metres and horizontal in kilometres).
M
Fig. 5. Location of the BandaSeis dataset regarding the well-tie several 2D seismic lines with well
data. Underlying base map: Global Multi-Resolution Topography (GMRT) synthesis (Ryan et al.,
2009).
ED
Fig. 6. NW-SE seismic 2D line opposite Timor Plateau (see figure 2) with geological interpretation
(inset map for location indicated by thicker pink line). This image shows clearly the process of fault
T
decapitation. The previous step is normal fault bending (green coloured fault) and finally as the
EP
bending process continues normal faults are decapitated (orange coloured fault) by thrusting (red
thicker coloured thrust) when arriving to the toe thrust wedge (vertical scale in metres and
horizontal in kilometres).
CC
Fig. 7. NW-SE seismic 2D line opposite Timor Plateau (see figure 2) with the geological interpretation
(inset map for location by the thicker pink line). Seismic image neatly reveals the collisional
A
processes at the orogenic toe-thrust with the NW-dipping normal fault is partially decapitated and
inverted (vertical scale in metres and horizontal in kilometres).
Fig. 8. NW-SE seismic 2D line opposite the Nunkolo hanging-wall (see figure 2) with the geological
interpretation (inset map for location by the thicker pink line). Example of Triassic detachment.
Normal faults bending, fault-propagation folds arriving to the toe-thrust wedge (vertical scale in
metres and horizontal in kilometres).
Fig. 9. Composite strike seismic line and interpretation along 750 km of Timor Trough (inset map for
location by the thicker pink line). This regional cross-section shows the main geological features
discussed in this paper: West Timor and Cova-Lima sub-basins accommodated by the Nunkolo and
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
Manatuto strike-slips faults as well as Timor Plateau (vertical scale in metres and horizontal in
kilometres).
Fig. 10. NW-SE seismic 2D line showing the Latest Callovian or breakup unconformity (Inset map for
location by the pink thicker line). Significant to point out the thickness variation of the light blue
horizon (Latest Callovian) as well as the thin Cretaceous succession veneering the break-up
unconformity. Also, this image provides a good example of Miocene geometries (clinoforms) and
bioherms (vertical scale in metres and horizontal in kilometres).
Fig. 11. NW-SE seismic 2D line opposite Timor Plateau (see figure 2) precisely uncovering how the
PT
collisional stress field is transmitted southwardly almost 10 km from the orogenic toe-thrust front
(inset map for location by the thicker pink line). The reactivation and inversion of normal faults gives
rise to a rugose bathymetry (vertical scale in metres and horizontal in kilometres).
RI
SC
U
N
A
M
T ED
EP
CC
A
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
PT
RI
SC
U
N
A
M
T ED
EP
CC
A
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
PT
RI
SC
U
N
A
M
T ED
EP
CC
A
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
PT
RI
SC
U
N
A
M
T ED
EP
CC
A
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
PT
RI
SC
U
N
A
M
T ED
EP
CC
A
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
PT
RI
SC
U
N
A
M
T ED
EP
CC
A
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
PT
RI
SC
U
N
A
M
T ED
EP
CC
A
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
PT
RI
SC
U
N
A
M
T ED
EP
CC
A
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
PT
RI
SC
U
N
A
M
T ED
EP
CC
A
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
PT
RI
SC
U
N
A
M
T ED
EP
CC
A
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
PT
RI
SC
U
N
A
M
T ED
EP
CC
A
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
PT
RI
SC
U
N
A
M
T ED
EP
CC
A
Downloaded from http://sp.lyellcollection.org/ at Macquarie University on August 30, 2019
PT
RI
SC
U
N
A
M
T ED
EP
CC
A