Sei sulla pagina 1di 19

Marine and Petroleum Geology 32 (2012) 76e94

Contents lists available at SciVerse ScienceDirect

Marine and Petroleum Geology


journal homepage: www.elsevier.com/locate/marpetgeo

Neogene development of the Savu Forearc Basin, Indonesia


James W.D. Rigg 1, Robert Hall*
SE Asia Research Group, Department of Earth Sciences, Royal Holloway University of London, Egham Hill, Egham, Surrey TW20 0EX, UK

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 12 July 2011
Received in revised form
23 October 2011
Accepted 4 November 2011
Available online 18 November 2011

The Savu Basin is situated in the SundaeBanda forearc where there is now a change from oceanic
subduction to continentearc collision. It records Middle Miocene initiation of subduction of the Banda
oceanic embayment, subsequent arc volcanism, and Pliocene to Recent collision of the Australian
continent and Banda forearc. New seismic data provide insights into its Neogene development. Four
major units can be recognised offshore and correlated with geology on land. Unit 1 is underlain by
continental crust and Cretaceous-Paleogene arc rocks capped by Oligocene to Lower Miocene shallow
water carbonates. Subduction rollback-induced extension in the Middle Miocene caused subsidence to
depths of several kilometres. Units 2 to 4 include Middle Miocene to Pliocene arc-derived volcaniclastic
turbidites and deep water carbonates that thin northwards into the basin, and Pliocene to Recent
sediments that thicken northwards due to slumping from the south. Apart from tilting and slumping the
Savu Basin is little deformed and there are no thrusts within it. Only at its southern edge is there
signicant deformation close to the Savu and Roti Thrusts, which do not form part of a single throughgoing thrust zone. There are two ridges at the south side of the basin. The Sumba Ridge was elevated as
Australian margin continental crust underthrust the forearc to form a broad exure tilting older units
and causing debris ows and turbidites to ow northwards into the basin. The SavueRoti Ridge is
composed of a pre-collision Banda forearc accretionary complex and Australian margin sedimentary
cover. It is partly bounded by north-vergent and south-vergent thrust zones and has risen from depths of
more than 2 km since 2 Ma. Most of the basin is undeformed but remains a high-risk area for hydrocarbons although there are a few indications which may be prospective in shallower areas.
2011 Elsevier Ltd. All rights reserved.

Keywords:
Indonesia
Sunda arc
Banda arc
Seismic
Extension

1. Introduction
The Savu Basin is part of the forearc of the SundaeBanda
volcanic arc at the southern Eurasian margin (Fig. 1). It is located
just north of the position where there is a change from subduction
of Indian Ocean crust at the Java Trench in the west to collision
between the Australian continental margin and the Banda forearc
to the east.
There are hydrocarbon provinces in the area surrounding the
Savu Basin and current interest in exploration in this frontier
region. The productive oil and gas elds of the northern Browse
Basin and Bonaparte Gulf Basin are located in the Australian NW
Shelf southeast of the Savu Basin. These are situated within the
relatively undeformed rocks of the Australian continental margin,
which are juxtaposed to the north against the highly deformed

* Corresponding author. Tel.: 44 1784 443897; fax: 44 1784 434716.


E-mail address: robert.hall@es.rhul.ac.uk (R. Hall).
1
Now at: CGGVeritas, Crompton Way, Manor Royal Estate, Crawley, West Sussex
RH10 9QN, UK.
0264-8172/$ e see front matter 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.marpetgeo.2011.11.002

rocks of the Banda ArceAustralian collision complex, notably in


Timor. There is considerable exploration interest in the Banda
region (e.g. Charlton, 2004; Satyana and Purwaningsih, 2011),
where hydrocarbon seeps have long been known, but petroleum
systems are complicated by a number of factors, including uncertainty about the nature of the underlying crust, its complex tectonic
setting, and the consequences of Neogene deformation. The Savu
Basin has been very little explored although recently acquired
seismic data from the basin show a number of features which may
be associated with hydrocarbons, and seeps have been identied
from satellite mapping (Toothill and Lamb, 2009). The basin has
been the subject of speculation concerning its origin because of its
position and unusual shape, especially the way in which it widens
westwards towards the forearc island of Sumba (e.g. Chamalaun
et al., 1982; Audley-Charles, 1985; van der Werff, 1995a; Rutherford et al., 2001). The islands surrounding the basin (Sumba,
Flores, Wetar, Timor, Roti and Savu) have been the source of most
information for its interpretation.
The Savu Basin is bounded to the west by the island of Sumba
and a submarine ridge (the Sumba Ridge) that crosses the forearc
obliquely in an NWeSE direction. It narrows from its greatest width

J.W.D. Rigg, R. Hall / Marine and Petroleum Geology 32 (2012) 76e94

77

Figure 1. Location of the Savu Basin in SE Asia. Bathymetry simplied from Gebco (2003) with contours in metres. Red triangles are active and Quaternary volcanoes (Siebert and
Simkin, 2002). Box outlines area of Figure 2. (For interpretation of the references to colour in this gure legend, the reader is referred to the web version of this article.)

of c.200 km to approximately 20 km in the east north of Timor and


in the west to approximately 50 km north of Sumba (Fig. 2). To the
north is the active volcanic arc including the island of Flores which
passes east into an extinct sector of the Banda arc between Alor and
Wetar. Volcanic activity in this sector ceased by about 2e3 Ma
(Abbott and Chamalaun, 1981; Scotney et al., 2005; Herrington
et al., 2011). Although there are several volcanoes on Flores,
Wheller et al. (1987) noted that those in western Flores, north of
Sumba, may also be extinct although activity must have ceased
quite recently as several deeply eroded Quaternary stratovolcanoes remain. To the north of these extinct sectors there is
northward-directed thrusting in the backarc region north of Flores
and Wetar (Fig. 2) that has been interpreted as indicating subduction polarity reversal (e.g. Silver et al., 1983; Snyder et al., 1996).
However, north of the active arc segment in eastern Flores, in the
sector immediately north of the central Savu Basin, there is no
backarc thrusting (Silver et al., 1983).
At the eastern end of the Savu Basin the forearc narrows to less
than 50 km between the collision complex on Timor and the
inactive Banda volcanic arc islands of Alor and Wetar to the north.
To the south of the Savu Basin are the small islands of Savu and Roti
and to the south of them is a sinuous bathymetric trough that
connects the Timor Trough to the Java Trench. Debate continues
about the signicance of the trough (e.g. Fitch, 1970, 1972; AudleyCharles and Milsom, 1974; Fitch and Hamilton, 1974; Hamilton,
1979; Audley-Charles, 1986b, 2004; Harris, 1991; Lorenzo et al.,
1998; Tandon et al., 2000; Spakman and Hall, 2010) and the timing
of collision (e.g. Audley-Charles, 1986a, 2011; Hall and Wilson,
2000; Keep et al., 2003; Harris et al., 2009).
There has been little information from the Savu Basin itself. A
small number of marine geophysical cruises and hydrocarbon
exploration investigations crossed the basin in the 1970s and
1980s and acquired bathymetric, gravity, magnetic and seismic

data (e.g. Hamilton, 1979; Silver et al., 1983; Karig et al., 1987;
Fortuin et al., 1992; van der Werff et al., 1994; van der Werff,
1995a,b; Harris et al., 2009). For most of the seismic lines data
quality is relatively poor and penetration is typically 1 or 2 s TWT
at best. Toothill and Lamb (2009) illustrated seismic sections from
surveys in 2002 and 2007 which collected new gravity and seismic
data. CGGVeritas kindly provided the seismic lines acquired during
those surveys which provided the basis for this study of the Savu
Basin.
2. Dataset and previous stratigraphic studies
The seismic data (Fig. 2) consists of 32 2D seismic lines, of which
the longest is 535 km, covering the southern and western parts of
the Savu Basin, with a total area of 50,000 km2, which were
acquired and processed by CGGVeritas. The results of two surveys
were combined for this study. The rst dataset was acquired in
2002 and comprises 2740 km of long-offset 2D data to a depth of
12 s TWT. The second was acquired in 2007 and comprises 3000 km
of long-offset 2D data to a depth of 8 s TWT (Toothill and Lamb,
2009).
No wells have been drilled in the basin. The only well in the
study area was a dry hole drilled on north Savu in 1975 (Savu #1 of
Harris et al., 2009) with TD at 1227 m. This well reveals melanges
and repetition due to thrusting, but cannot be used to correlate into
the basin. Although there are no wells it is possible to correlate the
stratigraphy offshore with that on land in some of the surrounding
islands. To the north, no correlation with Flores is possible. Only
a few of the lines approach within 30e50 km of the south coast of
the island and most terminate more than 100 km south of Flores.
Regional bathymetry (Gebco, 2003; Sandwell and Smith, 2009),
and the few available older lines that do approach Flores (Hamilton,
1979; Karig et al., 1987), show there is an abrupt bathymetric

78

J.W.D. Rigg, R. Hall / Marine and Petroleum Geology 32 (2012) 76e94

Figure 2. Geographical features of Savu Basin and surrounding area. A. DEM of satellite gravity-derived bathymetry combined with SRTM topography (Sandwell and Smith, 2009).
SR is Sumba Ridge and SRR is Sumba-Roti Ridge. B. Contoured bathymetry in metres (Gebco, 2003) and location of seismic lines. Red triangles are active and Quaternary volcanoes.
Flores and Wetar Thrusts from Silver et al. (1983). Small red box north of Savu shows the area of the seismic grid of Harris et al. (2009). (For interpretation of the references to colour
in this gure legend, the reader is referred to the web version of this article.)

change south of the island with water depths rising from more than
3 km, at the northern ends of the new seismic lines, to almost sea
level over a distance of 10e15 km which we suggest is a major fault
(see below). However, the lines do provide good coverage close to
the island of Sumba, as well as Timor, Savu and Roti, although these
three islands are separated from the main basin by thrust zones. All
these islands emerged in the Pliocene or Pleistocene.
The stratigraphy and structure of Sumba has been well documented (e.g. van Bemmelen, 1949; Burollet and Salle, 1981, 1982;
von der Borch et al., 1983; Effendi and Apandi, 1993; Fortuin
et al., 1992, 1994, 1997), it is relatively undeformed, and the
seismic lines close to Sumba cross sequences that are also little
deformed. There are no major structures separating Sumba from
the sequences seen in the Savu Basin which makes it possible to
correlate from land to the offshore area. The oldest rocks reported
from Sumba (Roggeveen, 1929) are Jurassic sands and shales with
Inoceramus and ammonite fragments. Cretaceous marine tuffaceous sandstones and mudstones associated with basalt ows and
basaltic-andesitic agglomerates are intruded by a suite of igneous
rocks including granodiorites and gabbros. The older rocks are
unconformably overlain by Paleocene and Eocene volcanic rocks,
and non-marine to shallow marine conglomerates, sandstones, and
mudstones. Above these are widespread shallow water carbonate

deposits of Late Oligocene to Early Miocene age which are overlain


by thick deep marine volcaniclastic sedimentary rocks. The stratigraphic column (Fig. 3) shows the strata present in the eastern part
of Sumba (Fortuin et al., 1992, 1994) and correlation with the
seismic sequences identied in this study.
To the east of the Savu Basin metamorphic rocks and a sedimentary cover sequence derived from the Banda forearc were
emplaced as thrust sheets on Timor during the late Neogene and
have become known as the Banda Terrane (Audley-Charles and
Harris, 1990; Harris, 1991; Audley-Charles, 2011). The metamorphic rocks are assigned to the Mutis Metamorphic Complex in
West Timor (Earle, 1981; Brown and Earle, 1983) and correlated
with the Lolotoi Metamorphic Complex of East Timor (AudleyCharles, 1968; Barber and Audley-Charles, 1976) which Standley
and Harris (2009) have shown includes an Upper Cretaceous
sedimentary protolith and records an Eocene metamorphic event
interpreted to have occurred in the forearc.
Audley-Charles (1985) summarised the stratigraphy of the
Banda Terrane cover sequences in West Timor based on
earlier studies by Tappenbeck (1940), Audley-Charles and Carter
(1972), Audley-Charles et al. (1974), Haile et al. (1979), Rosidi et al.
(1996), and Earle (1981). The Palelo Group includes Upper
Jurassic agglomerates and tuffs, Lower Cretaceous cherts and

J.W.D. Rigg, R. Hall / Marine and Petroleum Geology 32 (2012) 76e94

79

Figure 3. Seismic stratigraphy of the Savu Basin and correlation with stratigraphy of east Sumba (Fortuin et al., 1992). The colours shown next to the seismic section for the four
units are those used on gures of other seismic proles. The vertical scale of seismic sections is two way travel time (TWT) in seconds. (For interpretation of the references to colour
in this gure legend, the reader is referred to the web version of this article.)

pelagic limestones, and a Paleocene sequence of volcanics and


volcaniclastics, resting with an angular unconformity on the
Cretaceous rocks. Eocene shallow water limestones with an Asian
character (Lunt, 2003) locally overlie the Mutis Metamorphic
Complex in West Timor with faulted and possible unconformable
contacts and there are widespread Upper Oligocene-Lower
Miocene shallow water limestones assigned to the Cablac Limestone that rest unconformably on the Mutis Metamorphic Complex,
locally with a basal conglomerate. Mio-Pliocene volcanic rocks,
including tuffs, are the youngest parts of the Banda Terrane.
Audley-Charles (1985) drew attention to the similarity of the West
Timor and Sumba sequences despite a separation of more than
400 km, and noted that Miocene volcanism seemed more important in Sumba than Timor.
To the south of the Savu Basin studies on and close to Savu and
Roti (Rosidi et al., 1996; Harris et al., 2009; Roosmawati and Harris,
2009) also provide valuable information that has aided interpretation of the offshore data, although both these islands, and the
offshore areas close to the islands are much more deformed than
the areas to the north in the basin itself, and it is not possible to
correlate from them into the basin. Marine geophysical investigations combined with gravity and tomographic modelling have
recently been used to interpret the deep crustal structure along
a regional transect crossing the central part of the Savu Basin
(Shulgin et al., 2009). The combination of information from
previous studies with the new seismic data set offshore provided
the basis for our interpretation of the stratigraphy offshore and the
history of the forearc.
3. Offshore and onshore stratigraphy
Four major units can be mapped through the basin (Figs. 4e6)
and most of the adjacent offshore areas and can be correlated with
sequences on land. The units have been described in Rigg and Hall
(2011) and only key points are repeated here.

3.1. Unit 1
Unit 1 is the deepest unit with a package of bright reectors up
to 0.4 s TWT thick at its top that can be mapped throughout the
entire area as a continuous feature despite offsets by extensional
faults (Fig. 4). We suggest this bright package correlates with
Eocene to Lower Miocene Nummulites and shallow water limestones observed on Sumba (von der Borch et al., 1983; Fortuin et al.,
1992, 1997; van der Werff et al., 1994) and Timor (Audley-Charles,
1985). The eld relationships described by Fortuin et al. (1992,
1994) from Sumba with tilted fault blocks capped by discontinuous carbonates, all overlain unconformably by Middle Miocene
and younger strata, closely resemble features seen in the seismic
lines. On Sumba thin conglomerates containing older shallow
water limestone clasts mark a probable Late Burdigalian unconformity that passes up into chalks and marls (Fig. 3). Fortuin et al.
(1994) suggest these conglomerates correspond to breakup of
a carbonate platform and rapid subsidence in the early Middle
Miocene. If our correlation is correct, and the reectors that can be
traced into the deepest parts of the basin are Lower Miocene and
older shallow water carbonates, the Savu Basin is very unlikely to
be underlain by young oceanic crust (Fleury et al., 2009) or be
a Paleogene marginal basin (Harris, 2003).
Unit 1 can on some lines be split into two. In some places there is
no clear boundary between the upper and lower parts whereas
elsewhere there is a sharp boundary. The upper part contains
localised sub-parallel reectors, and the faults that cut the top of
Unit 1 cannot be traced downwards, whereas the lower part is
characterised by laterally discontinuous bedding which can be
traced for distances of up to 20 km. We suggest that the lower part of
Unit 1 can be correlated with Mesozoic rocks on land, and the upper
part is probably Paleogene. On Sumba there are shallow marine
Jurassic sediments (Roggeveen, 1929), Cretaceous marine siltstones
and sandstones (Burollet and Salle, 1981, 1982) interpreted as
submarine fan deposits (von der Borch et al., 1983) and arc rocks

80

J.W.D. Rigg, R. Hall / Marine and Petroleum Geology 32 (2012) 76e94

Figure 4. Seismic line from western part of Savu Basin showing relationships between Units 1 to 4. Half graben at base of Unit 2 are clear at this location. High angle normal faults in
Unit 3 are well imaged.

that include volcaniclastics cross-cut by intrusive igneous bodies


(Effendi and Apandi, 1993; Fortuin et al., 1994, 1997; Abdullah et al.,
2000). These are overlain unconformably by non-marine to shallow
marine Paleogene conglomerates, sandstones, and mudstones,
foraminiferal limestones, tuffs and ignimbrites (von der Borch et al.,
1983; Effendi and Apandi, 1993; Abdullah et al., 2000).
3.2. Unit 2
Above the distinctive reectors at the top of Unit 1 is Unit 2
(Figs. 4e6) with a thickness up to 1.6 s TWT which lls the depressions created by the extensional faulting and onlaps the faults. Its top
is a prominent unconformity that can be mapped throughout the
basin at the base of Unit 3. Unit 2 is variable in character. It commonly
lacks the well bedded character of the overlying Units 3 and 4.
Reectors are clearest where the unit is thin and are typically bright,
sub-parallel and discontinuous, whereas thicker parts of Unit 2 are
characterised by a transparent seismic character.
We suggest Unit 2 can be correlated with volcaniclastic turbidites on Sumba (Fortuin et al., 1992, 1994) above the Late Burdigalian unconformity. Locally well bedded parts of Unit 2 pass
laterally into unreective packages consistent with a transition
from bedded turbidites to rapid deposited disorganised volcaniclastic material. On land these rocks are poorly dated due to
reworking and absence of microfossils but are considered to
explosive island arc volcanism between 14.8 and 5.5 Ma that waned
during the Tortonian and revived during the Messinian (Fortuin
et al., 1994, 1997). Fortuin et al. (1997) suggested subsidence
below a carbonate compensation depth of 4e5 km during the

Middle Miocene in east Sumba and during the Late Miocene in


central Sumba based on dissolution of carbonate in some samples
but the presence of calcareous nannofossils in others.
3.3. Unit 3
Unit 3 is a very distinctive sequence characterised by closely
spaced strong reectors particularly on seismic lines crossing the
southern part of the basin. It has a maximum thickness of 1.8 s TWT
and is typically about 1 s thick. On the NWeSE-trending Sumba
Ridge Unit 3 is at lying and conformable on Unit 2 but to the north
of the ridge there are a series of northward-dipping planar packages. In many places Unit 3 appears to downlap onto an unconformity at its base (Fig. 4) and in the deeper parts of the basin it
rests locally on Unit 1, and Unit 2 is missing (Fig. 7). On the few
undeformed sections south of the Sumba Ridge a similar downlap
onto underlying units can be seen. Unit 3 thickens north from the
Sumba Ridge into the deep Savu Basin and further north becomes
thinner, suggesting it was derived from the south. On the Sumba
Ridge it is cut by many high angle small displacement normal faults,
with a spacing of 8e10 km.
There are a number of slumped packages within Unit 3, well
imaged by marked contrasts in reector character (Fig. 7). They
have a chaotic internal structure compared to adjacent well bedded
sequences and thicknesses commonly between 0.1 and 0.35 s TWT.
The thinnest slumps cannot be mapped between lines indicating
they have a lateral extent of less than a few kilometres, but larger
slumps can be mapped within an area of about 25 km radius. The
variation in the thickness and number of slumps contributes to

J.W.D. Rigg, R. Hall / Marine and Petroleum Geology 32 (2012) 76e94

81

Figure 5. In the centre of the Savu Basin many high angle normal faults cut Unit 1 with local small half graben at the base of Unit 2. Units 2 and 3 are thin. Unit 4 can be subdivided
into several sub-units with prominent unconformities between them, which separate phases of deposition and slumping.

Figure 6. Approximately north-south line crossing the Savu Basin. At the south end Unit 1 is uplifted and thrust northwards towards the basin and Units 2, 3 and 4 are largely
missing and interpreted to have been redeposited in the basin as Unit 4.

82

J.W.D. Rigg, R. Hall / Marine and Petroleum Geology 32 (2012) 76e94

Figure 7. At the western edge of the Savu Basin slumped packages within well bedded sediments of Unit 3 record downslope movements at the basin edge. Individual slump
packages occur at various levels within Unit 3, and are localised features, none of which can be correlated between lines. There is considerable variation in internal structure. The
upper slumped section in A has very well dened toe thrusts associated with folding at its distal end.

lateral thickness changes in Unit 3. They are most common at the


southern edges of the basin and indicate repeated failures at the
basin margins as Sumba, Savu and Roti were uplifted.
We interpret Unit 3 as equivalent to the Tortonian to Pliocene
foraminiferal chalks, marls, hemipelagic nannofossil oozes and
volcanogenic muds with some thickly bedded volcanic mass ow
deposits on Sumba (von der Borch et al., 1983; Fortuin et al., 1992,
1994, 1997). The fan of sediment was derived from the southwest
and deposited in the deeper area of the central Savu Basin, and can
be traced at least 75 km northwards from the present southern
edge of the basin. This implies the Sumba Ridge was a slightly
elevated feature at the south edge of the basin when the fan began
to develop. The fan prograded northwards and beds downlap onto
older units. Local onlap between layers resembles features interpreted on Sumba by Fortuin et al. (1997) to be the result of shifting
fan lobes which may contribute to thickness variations in Unit 3.
The present dip of beds (up to 5e6 ) is unlikely to relate to an
original depositional slope for ne grained sediments and we
suggest this apparent downlap north of the Sumba Ridge is the
result of uplift of the ridge and rotation of beds. This implies that
uplift of the ridge began during deposition of Unit 3, supported by
increasing abundance of slumps up section.
3.4. Unit 4
Unit 4 can be mapped throughout the area north of Sumba Ridge
(Figs. 4 to 8) but it is not possible to correlate across the ridge to the

south because Units 2 to 4 now thin towards the ridge and are
absent on it. Locally Unit 4 onlaps Unit 3. It shows considerable
variation in thickness, can be subdivided into several parts, and is
characterised by slumping. Unit 4 is interpreted to represent
slumping of Unit 3 (and possibly parts of Unit 2) northwards into
the basin as the Sumba Ridge was elevated with several phases of
slumping (Fig. 4) separated by minor unconformities. Slumped
packages were derived from the south and backstep southwards as
the basin was lled, facilitated by extensional faulting at the top of
the slope, preserved in Unit 3.
The slumps have many features similar to deep water debris
ows, from cohesive slumps to large mass transport complexes,
described by McGilvery and Cook (2003) and Bull et al. (2009). They
are characterised by a chaotic and transparent internal seismic
character and include headwall scarps, internal thrusts, folds and
ramps, and small pop-up structures. Local cohesive slumps, formed
by downslope creep, have well dened toe thrusts and associated
folding at their distal end. Large blocks are often present within
larger mass transport complexes which in places retain some
bedded character. Disturbed units are separated in places by well
bedded units suggesting episodes of major slope failure alternating
with smaller slumps or turbidite ows.
There are a number of ramp features in the upper part of Unit 4
(Fig. 8) with an average height of about 0.5 s TWT. These are slumps
with a well dened base, which detached at a bedding plane, in
which the basal thrust ramped up against underlying sediment at
the distal end. The front walls of these features are often steep,

J.W.D. Rigg, R. Hall / Marine and Petroleum Geology 32 (2012) 76e94

83

Figure 8. Ramp features in Units 3 and 4 are associated with the transport of material from uplifted areas into the Savu Basin. A and B show variation in size. They cut both up and
down, truncate surrounding reectors, and are characterised by a chaotic internal structure with steep front walls.

ramping up against, and truncating strata already in place. Tilting of


strata has facilitated the downcutting associated with the ramps.
The generation of such ramp structures has been discussed by Bull
et al. (2009) based on 3D seismic data, and their development is
thought to reect localised variations in stress conditions for shear
failure. Downcutting may also reect a component of gouging
during transport and deposition (Posamentier and Kolla, 2003).
The thickness of Units 3 and 4 varies signicantly from east to
west. Unit 3 is thin in the SE close to Timor and Roti (up to 0.25 s
TWT) and Unit 4 is thick (about 1 s TWT). In the SW, close to Sumba,
the thicknesses of Units 3 and 4 are reversed. The number and
thickness of slumps associated with Unit 4 increase in importance
towards the east. In the northern part of the basin, Unit 4 is better
bedded with at, parallel and bright reectors which are less
prominent at greater depths. These internger with the slumps.
The well bedded sequences could represent turbidite deposits at
the distal ends of debris ows or material transported from the
volcanic arc to the north. Interngering of bedded deposits with
slumps of Unit 4 favours a Flores origin. There are also some
slumped horizons at the very top of the seismic sections closest to
Flores suggesting deposition downslope from the volcanic arc.
Features similar to those of Unit 4 have been described from
Sumba and seismic lines crossing the Savu Basin (van Weering
et al., 1989a,b; Fortuin et al., 1992; van der Werff et al., 1994;
Roep and Fortuin, 1996) and were suggested to indicate simultaneous uplift and downslope transport of debris into the northern

part of the Savu Basin during the Early Pliocene. Unit 3 largely
represents the depositional phase of development of the fan and
thinning to the east in the Savu Basin reects a greater distance
from the sediment source to the SW. Unit 4 represents the next
stage of development involving uplift. Its distribution can be
assessed from the isochron map (Fig. 9) and the three lines which
cross the basin from east to west. Our new data show that the area
affected by slumping is at least 18,000 km2 and is much greater
than previously estimated. We suggest that the fan had two sources. Initially, material was derived from Timor and spread out
towards the northwest. In the western part of the basin, Unit 4 was
derived from the western Sumba Ridge and the southern end of
Sumba, and carried eastwards. There was little material carried
north from Sumba.
4. Deformation history
The principal structures mapped from our seismic dataset are
shown on Figure 10. In the major part of the basin sediments are
at-lying with normal faulting in the deeper parts of the section.
There are some differences from earlier studies. Thrusts shown in
the Savu Basin by van der Werff (1995a) are probably slumps, and
he shows a different orientation of structures SE of Sumba that we
attribute to the much wider spacing and poorer quality of the older
seismic data. On regional maps major thrust zones are often traced
offshore across the Savu Basin, typically from Timor. For example,

84

J.W.D. Rigg, R. Hall / Marine and Petroleum Geology 32 (2012) 76e94

Figure 9. Isochron map of Unit 4 shows thickening towards the north suggesting that most of the material, including many slumped packages, is derived from the northern margin
of Sumba and Savu, with material owing towards the northeast. Unit 4 thickens away from the base of slope at the southern edge of the basin.

the Savu Thrust is exposed on land at the northern margin of Savu


Island (Harris et al., 2009) and is sometimes shown as major
continuous feature crossing the south side of the Savu Basin (e.g.
Audley-Charles, 1985, 2004; Harris, 2006; Shulgin et al., 2009). In
fact, there is no continuous thrust zone crossing it and most of the
Savu Basin is little deformed apart from tilting and slumping,
except close to the islands of Savu and Roti where there are zones of
thrusting which can be broadly described as the Savu Thrust and
the Roti Thrust.

13 km across, and has a small 100 m deep crater-like centre.


Internally, the uppermost 0.3 s is well stratied feature above
a section of poor quality seismic suggesting movement of material
from beneath. The feature resembles mud volcanoes found in areas
of rapid sedimentation where there is active hydrocarbon generation (Dimitrov, 2002). However, since the feature is cut by only one
line its shape is uncertain. It could be part of a submarine turbidite
canyon system with levees which is consistent with the well
stratied upper part of the feature.

4.1. Sea oor features

4.2. Normal faults in the Savu Basin

The irregular topography of the sea oor is an indication of local


active deformation. On the slopes descending from the Savu Ridge,
close to the basin oor, this is probably due to recent downslope
movements of material. This is clear on lines perpendicular to the
slope but is more difcult to prove on lines sub-parallel to strike.
These show small scarps with heights up to 100 m and abrupt
termination of bedding at these features on the seabed, but slumps
or mass transport complexes are difcult to identify.
Narrow canyons are seen on seismic lines that run across the top
of the slope at the southwestern edge of the Savu Basin close to
Sumba (Fig. 11). McGilvery and Cook (2003) suggest that such
canyons, the majority of which are steep walled, developed by
excavation through gravity mass wasting and can act as local
sediment conduits. They are clearly very young as they incise well
bedded horizontal strata at the top of Unit 4. It is probable that the
canyons are erosional products of material carried from Sumba
where in the western part of the island there are deeply incised
steep-sided valleys, running mainly northeast, from the highest
mountains of the island, implying young uplift on land.
Two areas of similar prominent sea oor topography are present
in the northern part of the basin (Fig. 12). One shown on Figure 12B
rises approximately 300 m from the sea oor in the deepest part of
the basin south of Flores. It has a low gradient prole, is roughly

Many large extensional faults cut the bright reectors at the top
of Unit 1 and created a series of full and half graben, which we
interpret to have formed in the Middle Miocene based on correlation with limestones on Sumba. Faults at the northern edge of the
Sumba Ridge, are typically 15 km apart, and although the extension
direction cannot be precisely determined because of the wide
spacing of seismic lines, it is broadly north-south. There are
numerous normal faults further north in the Savu Basin (Fig. 10).
Extension on these faults contributed to the initial subsidence of
Unit 1 which is now at depths of 2 s on the Sumba Ridge to 7 s TWT
in the deepest parts of the Savu Basin. The faults have not been
inverted, with the exception of some close to the Savu and Roti
Thrusts. The main Savu Thrust does appear to have reactivated at
the position of a previous normal fault (Fig. 13). The relatively small
number of shallow depth (<50 km) earthquake solutions (Das,
2004; GCMT, 2011) within the basin show normal and strike-slip
faulting in the basin but no indication of thrusting.
4.3. Savu Thrust
The Savu Thrust is not a single thrust but a zone of thrusting
(Figs. 10 and 13), as shown by Harris et al. (2009) on land and
offshore of northern Savu. We have not attempted to map the

J.W.D. Rigg, R. Hall / Marine and Petroleum Geology 32 (2012) 76e94

85

Figure 10. Summary structure map with the principal features identied in the area of study. The coloured shaded area north of the Java Trench and Timor Trough is, from west to
east, the transition from the Sunda forearc, which passes east into a deformed zone including forearc and Australian sedimentary cover forming the SavueRoti Ridge, into the
arcecontinent collision zone of Timor. The ne dotted black line north of Timor is the inferred northern limit of Australian continental basement. The dashed red line below Sumba
is the inferred northern and western limits of the underthrust Scott Plateau. Most of the Savu Basin is undeformed, except by slumping, and major structures in the basin are normal
faults. Thrusts are observed only close to Savu and Roti. The heavy dashed line along the north side of the Savu Basin is suggested to mark a steep structure with a large vertical
offset which could be a thrust or normal fault and is named the Alor-Flores Fault. (For interpretation of the references to colour in this gure legend, the reader is referred to the web
version of this article.)

thrusts in detail as we have only a few widely spaced lines that


cross the closely spaced grid of 1975 seismic proles in the small
area mapped by Harris et al. (2009). The older seismic proles
image the upper part of the section (<1 s TWT), with one well that
penetrates to 1227 m TD, from which Harris et al. (2009) interpret
south-dipping and north-dipping thrusts to a depth of 5 km.
The new seismic lines allow interpretation to about 1.5 s TWT
below the sea oor in the structurally complex area close to the
thrusts. They show that the thrust zone dies out to the east and
west within about 50 km of Savu. It can be traced from the west
through north Savu and further east is offset to the south by
roughly 5 km. The total displacement diminishes to the east and
west. We could not identify any north-dipping thrusts near Savu,
although they are present much further south on the south side of
the SavueRoti Ridge (Figs. 2 and 10). All the thrusts dip south and
the largest has a displacement of between 0.4 and 1.3 s TWT.
Immediately in front of the largest thrust (Fig. 13) there is a footwall
syncline in Unit 3 and behind the main thrust, faults and associated
fold limbs are progressively steepened. Some thrusts reach the
seabed whereas others are blind but their geometry is obscured by
poor seismic quality in the uplifted areas. There are numerous small
inverted normal faults, both behind and in front of the main thrust,
which also dip south.
There are several episodes of deformation associated with the
thrust zone, marked by prominent unconformities in deformed
Unit 3. The rst major phase of movement folded deeper parts of
Unit 3 and folds are onlapped by younger reectors. There are
numerous subsequent subtle onlaps of reectors near the top of
Unit 3, associated with syn-deformation thickening away from the
main fault.

4.4. Roti Thrust


The Roti Thrust (Figs. 10 and 14) is about 25 km south of the
eastern strand of the Savu Thrust and is located offshore 20 km
north of Roti with a NNE-SSW orientation parallel to the coast for
almost the entire 70 km length of the island. It has a similar character to the Savu Thrust, being a zone of thrusting, with thrusts that
dip southwards and die out to the east and west. The thrust zone is
wider but the maximum displacement on the most important
thrusts is less than in the Savu thrust zone. The largest thrust has
a displacement of about 0.5 s TWT. There are some inverted normal
faults in front of the main thrust. The net displacement is greatest in
the centre of the thrust zone and appears comparable to the Savu
thrust zone. Like the Savu Thrust, several phases of deformation can
be identied by mapping of minor unconformities in the deformed
zone with onlap and thickening of packages in front of the main
thrust. As noted above, the Roti and Savu thrust zones do not link
together.
4.5. SE-directed thrusting SE of Sumba
In the area southeast of Sumba, on the south side of the Sumba
Ridge, there is a zone of thrusting roughly 60 km across that widens
towards the Lombok Basin to the west. There are a number of
anticlines (Fig. 15) well imaged on the seismic lines which are
1000e1400 m across associated with numerous NW-dipping thrust
faults (Fig. 10). The orientation of faults and fold axes is difcult to
determine because of the wide spacing of seismic lines in this area
but they are clearly oblique to the Sumba Ridge and we estimate
a trend of about 030 . We interpret these to be structures formed in

86

J.W.D. Rigg, R. Hall / Marine and Petroleum Geology 32 (2012) 76e94

Figure 11. Canyons on the seabed east of Sumba are suggested to be a result of erosion by turbidity currents from the direction of Sumba. The tilted fault blocks at the base of Unit 2,
present throughout the basin, are particularly well imaged. In this area there is less contrast between Units 3 and 4 than elsewhere.

the former accretionary complex north of, but close to, the former
Banda Trench.
The thrusts and folds are overlain unconformably by almost 1 s
TWT of sediments which dip very gently northwards. It is not
possible to correlate these with units in the Savu Basin as there are
no lines that cross from SW to NE of the Sumba Ridge but they
resemble Unit 3 and were probably derived from the NW, thin to
the southeast, and have been tilted by the rise of the SavueRoti
Ridge.

4.6. SE-directed thrusting of the SavueRoti Ridge


The SavueRoti Ridge is a broad zone of largely seismically
opaque, apparently highly deformed material. Within the ridge
are narrow zones in the upper 1 s TWT of the seismic sections with
NEeSW striking thrusts that dip northwards and reach the
seabed. There are two particularly clear zones of deformation
(Fig. 10). The majority of faults within the northern, slightly
broader, zone dip north, although some smaller antithetic faults
can also be seen. The southern zone is characterised by small
northward-dipping thrusts (Fig. 16), which become progressively
rotated and steepened towards the north. Between these two
zones some faults and reectors can be seen in places. Deformation in the SavueRoti Ridge may be active today and thrusts are
evidently younger than those close to Sumba since they are not
overlain by sediment.

4.7. Alor-Flores Fault


The northernmost part of the Savu Basin is not crossed by
any of the seismic lines of this study. There are also no seismic
lines by previous surveys that approach the coast of Flores,
undoubtedly because of the rapid change in water depth at the
northern edge of the Savu Basin. Here (Fig. 2) there is a spectacular slope with gradients of 15e20 that descends from the
Flores coast over a distance of 10e15 km to water depths of
more than 3 km (Hamilton, 1979; Karig et al., 1987) which is
also evident in global bathymetry data (Gebco, 2003; Sandwell
and Smith, 2009). The gradient is even greater if the land
morphology is included as volcanoes very close to the coast on
the south side of Flores have elevations of more than 1 km. Not
only is this slope very steep, it is quite unlike any feature in
any volcanic arc elsewhere in Indonesia in being so close to the
active arc itself.
Rutherford et al. (2001) speculated that there was a dextral
strike-slip fault parallel to the southern coastline of Flores dening
the edge of a former west-moving Sumba block, but inactive since
about 7 Ma. We did not nd most of the structures postulated by
them (e.g. Rutherford et al., 2001, their Fig. 5) in the basin and
surrounding region and our interpretation of timing of events is
very different. However, they are undoubtedly correct in identifying a fault along the north side of the basin but we suggest this is
an active fault with an important dip-slip component which we
name the Alor-Flores Fault (Fig. 10). No seismic lines cross the fault

J.W.D. Rigg, R. Hall / Marine and Petroleum Geology 32 (2012) 76e94

87

Figure 12. A. Basin oor structure close to Flores. There is some bedding, which suggests the end of a channel with levees sourced from the volcanic arc or a relatively low viscosity
mud volcano suggested by deeper disturbed section. B and C. Zones of disrupted bedding in the upper part of Unit 4 are suggested to be due to gas escape linked to minor faults.
Small offset faults are not uncommon at this depth but are normally sharp, as seen on the left of C, and not associated with such disruption.

but the steep slope and free air gravity gradient are illustrated by
Rutherford et al. (2001) and there is a concentration of earthquake
epicentres (Engdahl et al., 1998) close to the postulated fault. There
are only three earthquake solutions on the line of the fault (GCMT,
2011); assuming an ENE fault trend the easternmost is sinistral, the
second is a normal fault with a sinistral component, and the
westernmost is normal.
On the north side of the volcanic arc are the two active northdirected Flores and Wetar Thrusts (Fig. 2). These do not link to
form a single thrust zone and are separated by a zone of strike-slip
faulting on the north side of Flores (McCaffrey, 1988; Das, 2004;
GCMT, 2011) in the sector north of our interpreted Alor-Flores
Fault. It is not coincidental that the Flores and Wetar Thrusts are
north of the two rectilinear continental promontories of East Timor
and the Scott Plateau (see below), and in the sectors where volcanic
activity has ceased (Wheller et al., 1987). The Alor-Flores Fault could
be an extensional response to loading and tilting of the forearc
between the promontories where collision has caused volcanic
activity to terminate. It could alternatively be a steep thrust but
with the opposite vergence to the Flores and Wetar Thrusts, and
van der Werff (1995a) shows south-directed thrusts at the northern
end of two seismic lines close to Flores. If this is correct, the Flores
and Wetar Thrusts are unlikely to represent the early stages of
subduction polarity reversal, and the thrusts simply represent

shallow crustal features accommodating contraction across the


entire arc.
5. Discussion: Cenozoic evolution
The shape of the Savu Basin and the unusual way in which
Sumba crosses the forearc have been the cause of controversy and
speculation over many years (e.g. Hamilton, 1977, 1979; AudleyCharles, 1985; van der Werff, 1995a,b; Rutherford et al., 2001;
Fleury et al., 2009). It is now accepted that Sumba was part of the SE
Asian continental margin since the Cretaceous, and the Savu Basin
was part of the SundaeBanda forearc. Many authors have recognised the importance of trench retreat or rollback as an important
feature of Banda Arc development (e.g. Hamilton, 1976, 1979;
Harris, 1992, 2003, 2006; Hall, 1996, 2002; Charlton, 2000;
Milsom, 2001; Hinschberger et al., 2005). However, when the
Savu Basin reached its present depth, the nature of the underlying
basement, and how these relate to Banda rollback remain uncertain. We have proposed elsewhere (Rigg and Hall, 2011) a tectonic
interpretation of the Savu Basin that builds on previous studies of
the region and incorporates our new observations in the context of
a subduction rollback model which includes detailed reconstructions of the whole Banda Arc (Spakman and Hall, 2010). We do not
repeat the interpretation here but discuss below key points based

88

J.W.D. Rigg, R. Hall / Marine and Petroleum Geology 32 (2012) 76e94

Figure 13. Seismic line crossing the centre of the Savu Thrust north of Savu, where there is the maximum amount of displacement and folding in front of the thrust. Deformed material
south of the thrust includes Australian margin rocks and sediments from the southern margin of the Savu Basin and is separated from less deformed Units 1 to 3 in the basin.

on the new seismic data that are important in the late Cenozoic
development of the Savu Basin.
5.1. Basement
The Sumba margin was the site of arc volcanism during the
latest Cretaceous, Paleocene and Eocene (Abdullah et al., 2000;
Lytwyn et al., 2001). A continental basement for the Sumba
region was favoured by many early workers (e.g. Hamilton, 1977;
Bowin et al., 1980; von der Borch et al., 1983; van der Werff et al.,
1994) although it was uncertain if this had come from SE Asia or
Australia. Palaeomagnetic evidence indicates Sumba was part of the
Sunda margin by the Late Cretaceous (Wensink, 1994, 1997) and has
not moved signicantly relative to SE Asia since the Early Miocene
(Wensink & van Bergen, 1995). Rutherford et al. (2001) suggested
the Cretaceous arc was built on oceanic crust and Harris (2006)
described it as a continent-fringing arc mounted on oceanic basement. In contrast, van der Werff (1995a) interpreted a strong Savu
Basin to be underlain by rigid crust of continental origin. We
suggest that the region from SW Sulawesi to West Timor including
Sumba and the Savu Basin is underlain by Australian continental
crust that was accreted to SE Asia in the mid Cretaceous (Wakita
et al., 1996; Parkinson et al., 1998; Hall et al., 2009; Hall, 2011) on

which later arcs were constructed. This suggestion is supported by


crustal thicknesses and densities (Shulgin et al., 2009) and by
Jurassic ammonites and bivalves reported from South Sulawesi
(Sukamto and Westermann, 1993) and Sumba (Roggeveen, 1929).
Whatever the nature of the deep crust it is generally agreed that
volcanic activity ceased in the Eocene when the region must have
been close to sea level or emergent. There have been a number of
different suggestions for the development of the southeastern part
of the Sunda margin since the Eocene. Fleury et al. (2009) suggested
that the intact internal structure of the Savu Basin is due to the
rigidity of its oceanic basement. Harris (2006) interpreted Sumba
and the Sumba Ridge as remnants of a continental arc terrane
fragmented and dispersed by Oligocene extension, and Harris
(2003) showed the Savu Basin as a marginal basin formed at
25e35 Ma, appearing to imply that it is underlain by oceanic crust,
and certainly that it subsided in the Oligocene. In an earlier paper,
Harris (1992) suggested the eastern Savu Basin was underlain by
a supra-subduction zone ophiolite formed since about 6 Ma during
trench retreat which would suggest it subsided in the Late Miocene.
Although the present active volcanic arc is north of Sumba,
Fortuin et al. (1994) suggested that Sumba and the Savu Basin were
in a backarc position in the Middle Miocene and moved to a forearc
location since the Late Miocene. They did not identify a mechanism

J.W.D. Rigg, R. Hall / Marine and Petroleum Geology 32 (2012) 76e94

89

Figure 14. The Roti Thrust resembles the Savu Thrust but the zone of deformation extends further north than in the Savu area. Prominent unconformities within deformed material
are evident. It is not possible to correlate strata at this location with those further to the west near the Savu Thrust.

but this could be explained by oceanic spreading in the Savu Basin


associated with a movement of the arc, or by a jump of the arc from
south to north forming an interarc basin (Charlton, 2000) with
crust of an uncertain nature. This would imply the basin subsided
from the Middle Miocene.
We interpret Unit 1 in the Savu Basin to represent the Late
Cretaceous to Early Miocene interval and correlate the distinctive
reectors at the top of Unit 1 with Upper Oligocene-Lower Miocene
shallow marine limestones described on Sumba and Timor. This
horizon can be traced across the whole of the area covered by our
data implying that it was close to sea level in the Early Miocene.
This, and the structure of Unit 1, rules out suggestions that the
basement is oceanic crust. It also conicts with the suggestion of
Harris (2003, 2006) that there was major extension and formation
of marginal basins within the SE Asian forearc during the Oligocene
following collapse of the Sulawesi orogen. There may have been
oceanic crust further east in the Banda forearc, now preserved in
the Banda Terrane on Timor, but the area between West Timor and
Sumba including most of the Savu Basin did not subside before the
Middle Miocene and does not include oceanic crust.
5.2. Subsidence and volcanism
Subsidence on Sumba began in the Middle Miocene with
breakup of a carbonate platform leading to rapid subsidence from
about 16 Ma (Fortuin et al., 1994). This was an effect of the beginning of hinge rollback into the Banda embayment that led to the
formation of the Banda Arc (Spakman and Hall, 2010). There was
a signicant inux of volcaniclastic turbidites that we correlate
with Unit 2 deposited in deep water on Sumba and in the Savu

Basin (Fortuin et al., 1992, 1994). Fortuin et al. (1997) suggested that
these deposits were derived from a volcanic arc south of Sumba and
form part of a fan that prograded northwards. This arc should be
somewhere beneath the Sumba-Roti Ridge before it moved to its
present position in Flores north of the Savu Basin and Sumba.
However, volcanic activity is known on Flores from 16 Ma
(Hendaryono, 1998; Fleury et al., 2009) and there is no indication in
our seismic data of coarsening or thickening of beds to the south
consistent with another arc south of Sumba. We propose that the
arc was always in its present position north of the Savu Basin and
volcaniclastic debris from west Flores and Lombok was transported
ESE through the Lombok Basin and then turned northward into the
Savu Basin. Sediment movement was broadly parallel to the WNWtrending Sumba Ridge that followed a deep basement trend in the
underlying Australian continental basement that had been accreted
in the mid Cretaceous (Hall et al., 2009). The top of the ridge in
West Sumba was at very shallow depths through the Neogene but
became much deeper to the eastat which point sediment was able
to turn northwards. NE-SW orientated faults in Sumba, which
allowed East Sumba to subside to depths of 4e5 km (Fortuin et al.,
1994, 1997) would have opened a passage into the Savu Basin from
the SW.
In the Savu Basin extensional faults have a broadly E-W trend,
curving from WNW in the west to ENE in the east. We suggest these
extensional faults in the basin formed in response to subduction
and rollback into the Banda embayment which began after 16 Ma
(Hall, 2002, 2009, 2011; Spakman and Hall, 2010). Offshore to the
north of Sumba, faults that can be correlated between seismic lines
appear to be oriented NeS, which we interpret to be due to alongarc extension.

90

J.W.D. Rigg, R. Hall / Marine and Petroleum Geology 32 (2012) 76e94

Figure 15. SE-vergent folds and thrusts SE of Sumba. The deformation in this area is suggested to be related to subduction at the former Banda Trench and predate arcecontinent
collision. Folds and thrusts are locally overlain by 1 s TWT of sediment. Anticlines contain some at spots which cut across stratigraphy and may indicate the presence of previous, or
in place, hydrocarbons.

5.3. Strike-slip faulting and thrusting


There is no obvious evidence for the large strike-slip faults
shown to cross or bound the basin by some authors (e.g. SoeriaAtmadja et al., 1998; Rutherford et al., 2001) and surmised to
have produced thrusting SW of Sumba. Previously, in the absence of
offshore data, several authors suggested that north-directed
thrusting north of Timor could be connected to surmised offshore
structures that cross the Savu Basin or join the Roti and Savu
Thrusts (e.g. Audley-Charles, 1985; Roosmawati and Harris, 2009;
Shulgin et al., 2009). Our seismic data show that the Roti and
Savu Thrusts are discontinuous and there are no major thrusts in

the deeper part of the basin. In the central Savu Basin the only signs
of contractional deformation are slumps in Unit 4, and the seabed is
at or gently sloping.
The oldest thrusts found in this study are SE of Sumba and north
of the SavueRoti Ridge (Figs. 10 and 15). They have an ENEeWSW
trend and are now covered by up to 1 s TWT of sediment resembling Unit 3. We interpret them to have formed in the accretionary
complex close to the former Banda Trench (Breen et al., 1986).
The Sumba Ridge was elevated during the collision process
(Figs. 6 and 13) after the rst arrival of Australian continental
margin at the Banda Trench south of Sumba (Harris, 1991; van der
Werff, 1995a,b). Bedding of Unit 3 dips at up to 5e6 as a result of

Figure 16. The southern side of the deformed complex south of the SavueRoti Ridge shows SE-directed thrusting and folding. The thrust faults are well imaged at the southern end
of the uplifted area, and become progressively steeper towards the NW. As thrusts become steeper, internal deformation also increases, which results in a poorly imaged section.

J.W.D. Rigg, R. Hall / Marine and Petroleum Geology 32 (2012) 76e94

91

Figure 17. Localised discontinuous high amplitude reectors (outlined) in the upper part of Unit 4. Many, but not all, are close to faults, and there is an unusual abundance of them
in this area. They may indicate the presence of shallow gas.

uplift of the ridge and rotation of beds and there is an increasing


abundance of slumps up section. Slumps in Unit 4 are directed
basinward from Sumba and Timor. Deformation of Unit 3 would
have begun after 4 Ma based on correlation with Sumba where
there was uplift from depths of more than 5 km since then (Fortuin
et al., 1997). The present elevation of more than 1 km on Sumba
with carbonate reefs up to 500 m above sea level record uplift in the
last 1 million years (Pirazzoli et al., 1993). There is no observable
inversion of normal faults or thrusting associated with the uplift of
Sumba Island which appears to have risen as a broad upwarp.
The Savu and Roti Thrusts (Figs. 13 and 14) are much younger
than all other structures and are currently active zones of deformation that include multiple thrusts. The seismic lines show that
they are not lithosphere-scale features, nor major through-going
regional structures, as they die out rapidly to the east and west
and do not link up. Both zones include thrusts that emerge at the
sea oor, blind thrusts associated with deformation of the sequence
above, and folding of the seabed.
The accretionary complex that presumably formed north of the
Banda Trench during oceanic subduction has been overridden and/or
incorporated in the wedge of deformed material south of the Savu and
Roti Thrusts, except in the extreme SE where it is seen beneath at
lying sediments SE of Sumba. Unit 4 records slumping associated with
uplift of the SavueRoti Ridge and Roosmawati and Harris (2009) show
that signicant and rapid uplift from water depths of more than 2 km
to emergence of the islands of Savu and Roti began at about 2 Ma. The
broad zone of seismically opaque material with narrow thrust zones
would therefore have been deformed since then. Most of the ridge
must include material from both the Banda forearc and the Australian
continental margin, as also suggested by observations on land in Roti
and Savu (Harris et al., 2009; Roosmawati and Harris, 2009).
Thrusting south of Sumba and the Savu Basin has two causes:
deformation associated with subduction at the Java and former

Banda Trench, and collision of the Banda forearc with the Australian
margin which began at around 4 Ma (e.g. Audley-Charles, 1986a,
2011; Harris, 1991; Hall, 1996, 2002; Spakman & Hall, 2010). The
collision was also inuenced by the shape of the Australian margin
which had rectilinear offsets north of East Timor and the Scott
Plateau similar to those seen today at the Exmouth Plateau of the
NW Australian shelf. This is the major cause of the unusual shape of
the Savu Basin. Fortuin et al. (1994) noted that the Sumba region
departed from timeespace equivalence predictions for the western
Banda orogen of Harris (1991). Harris explained this as the result of
collision of the Scott Plateau whereas Fortuin et al. (1997) postulated a role for their surmised southern volcanic arc. We agree with
Harris (1991). The rst volcanic arcecontinent collision was north
of East Timor and deformation propagated west with time in Timor
(Harris, 1991). It then began independently in Sumba as the Scott
Plateau came into contact with the forearc. The shape of the margin
means that beneath Savu and Roti Australian continental crust has
only recently underthrust the leading edge of the Banda forearc
(Rigg and Hall, 2011).
6. Implications for hydrocarbon exploration
Forearc regions are generally not considered attractive targets
for hydrocarbons. Because there has been little exploration and
almost no drilling in the region around the Savu Basin the age and
nature of the basement and overlying units was not known. Toothill
and Lamb (2009) suggested possible hydrocarbon sources based on
comparison with Seram and Timor, and the pre-collisional rifted
margin of northern Australia. They interpreted the age of rifting at
the top of Unit 1 as Mesozoic whereas we suggest it is Middle
Miocene, and we therefore differ in suggestions of possible source
rocks for a petroleum system. We agree that the deep basement is
likely to be Australian continental crust with a structure inherited

92

J.W.D. Rigg, R. Hall / Marine and Petroleum Geology 32 (2012) 76e94

from the pre-rifted Australian margin but suggest that this


Australian basement arrived in the Cretaceous. Regional deep
seismic lines have indicated a thick pre-Cenozoic section (Emmet
et al., 2009; Granath et al., 2011) in the East Java Sea possibly
analogous to the Canning Basin based on tectonic reconstructions
(Hall et al., 2009; Hall, 2011) which could be present beneath
Sumba. This could provide a possible petroleum source although
much deeper than anything interpreted from seismic lines of this
study.
Unit 1 may include source rocks. Based on correlation with
Sumba the upper part is likely to include shales, mudstones, siltstones and tuffaceous sandstones (van Bemmelen, 1949; Burollet
and Salle, 1981, 1982; von der Borch et al., 1983; Effendi and
Apandi, 1993; Fortuin et al., 1994; van der Werff, 1995a) with
small Paleogene intrusions. Hamilton (1979) mentions lignites.
There is limited evidence for hydrocarbon generation in the basin
from these sources, or others. Toothill and Lamb (2009) drew
attention to satellite mapping of oil slicks on the surface of the Savu
Sea and the position of some of them above structures identied on
seismic lines. Steeply dipping beds at the sea oor or faults which
penetrate to the seabed can be directly linked to slicks on the
surface of the Savu Sea. The correlation of these potential conduits
with the slicks on the sea surface would further suggest the
generation of hydrocarbons at some level within the subsurface
(Toothill and Lamb, 2009).
Near the northern edge of the Savu Basin there are uid-escape
structures that could be gas chimneys (Figs. 12 and 17). They are
different from faults well imaged elsewhere at the same depth in
the section, and there is a change in seismic character in the
surrounding subsurface, not seen in faulted areas, which originates
and widens in a triangular fashion from a point at depth. They may
be associated with leakage from deeper gas accumulations rather
than breaching of shallow hydrate layers. If so, they would indicate
the presence of a generating source rock within the basin. Possible
mud volcanoes (Fig. 12) suggest overpressured zones which could
be associated with hydrocarbons although these also suggest the
risk of damage to seals.
In some areas just north of the northern coast of Sumba (Fig. 17)
numerous short (up to 500 m) localised, high amplitude reectors
are visible at depths of 2e3 s TWT. They could be microbial gas
deposits formed at shallow depths and buried due to high sedimentation rates (Floodgate and Judd, 1992). It is also possible that
these could have a thermogenic origin and have migrated from
depth. The small bright spots suggest trapping against the extensional faults that are present in Unit 3. The small and localised
nature of the accumulations may relate to the lack of a laterally
continuous sealing horizon, the lack of a suitable reservoir or the
signicant deformation experienced by Unit 3.
The relative lack of deformation throughout most of the Savu
Basin suggests that the integrity of a traps is not a major concern.
Possible reservoirs are carbonates at the top of Unit 1 and possibly,
but much less attractive, the volcaniclastic sandstones in Units 2 to
4. There may be stratigraphic traps formed by the numerous
onlapping surfaces. Rotated domino-style fault blocks could also
serve as potential hydrocarbon traps at depth. Inversion anticlines
have been identied as possible traps in Banda area by Charlton
(2004) and well imaged anticlines are located 35 km to the south
of Sumba (Fig. 15) within the Sumba Basin. Associated with these
are prominent at spots, which are between 600 and 700 m in
length. These features appear to be within a layer 0.15 s thick,
bounded on either side by bright horizons. These could be hydrocarbons or may represent a palaeo oil-water-contact, marking
diagenetic effects of previously in-place hydrocarbons. Overall,
there are possibilities for both hydrocarbon sources and reservoirs,
but on the basis of information available at present this is clearly

a high risk area. Much of the basin is undeformed but water depths
are greater than 3 km. The relatively shallow areas close to Sumba,
and the Sumba Ridge are the most accessible.
7. Conclusions
The Savu Basin is probably underlain by Australian continental
crust accreted to the SE Asian margin in the Cretaceous and was the
site of volcanic activity in the Late Cretaceous and Early Cenozoic.
From the Late Eocene to Early Miocene it was close to sea level and
carbonates were deposited over a widespread area, including
Sumba and the Savu Basin. It subsided rapidly to depths of several
kilometres in the late Middle Miocene in response to extension
induced by rollback at the Banda Trench as subduction propagated
east into the Banda Embayment. The extension was marked by
widespread normal faulting.
From the Middle Miocene a thick succession of volcaniclastic
turbidites was deposited in the basin. They entered the basin from
the SW but were probably derived from the volcanic arc to the
north of Sumba. Transport was inuenced by the Sumba Ridge in
the forearc which caused turbidity currents to ow rst ESE parallel
to the ridge, and then NE into the Savu Basin. The ridge is suggested
to be a feature that reects the deep structure of the Sunda margin.
The western part of the ridge remained a shallow bathymetric
feature during the Neogene although the eastern part subsided to
depths of several kilometres.
Apart from tilting and slumping the Savu Basin is little deformed
and there are no thrusts within it but its southern margin records
Pliocene to Recent collision of the Australian continent and the
forearc. Thick Middle Miocene to Pliocene sediments of the Savu
Basin thin northwards into the basin, whereas Pliocene to Recent
sediments thicken northwards due to slumping into the basin from
the south. Sumba and the Sumba Ridge was elevated as continental
crust of the Australian margin arrived at the Banda Trench and was
exed into a broad upwarp that tilted the volcaniclastic turbidite
sequence and later caused debris ows and turbidites to ow
northwards into the basin. Slumps came from both Sumba and
Timor.
Collision began earliest in the east in East Timor where the
collision complex is now almost in contact with the former volcanic
arc. Rectilinear offsets in the former Australian continent-ocean
boundary meant that forearc-continent contact occurred later in
the SavueRoti area at the south side of the basin. This is the reason
for NE-narrowing of the Savu Basin. In the west the northern edge
of the Scott Plateau is now beneath the Sumba Ridge, contributing
to the elevation of Sumba and extensional collapse seen on the
island.
SE of Sumba a small part of the original forearc accretionary
complex is preserved, now buried beneath almost 1 s of sediment.
The SavueRoti Ridge is composed partly of Banda forearc sediments and partly of Australian margin sedimentary cover. Most of
the pre-collisional accretionary complex has been incorporated in
the ridge south of the Savu and Roti Thrusts including the islands of
Savu and Roti which rose from depths of more than 2 km since 2 Ma
and emerged very recently. The deformed complex is partly
bounded to the north and south by north- and south-vergent
thrusts; the former trench is now deep beneath this complex and
the lithospheric faults do not emerge at the surface. To the south
the surface of this wedge dips south and is underlain by Australian
margin sedimentary cover.
On the north side of the Savu Basin is the volcanic arc and north
of the arc are the active north-directed Flores and Wetar Thrusts.
These are located north of the two rectilinear continental promontories of East Timor and the Scott Plateau in sectors where
volcanic activity has ceased. Between them on the south side of

J.W.D. Rigg, R. Hall / Marine and Petroleum Geology 32 (2012) 76e94

Flores is a major fault, very close to the volcanic arc, which is


suggested to be a normal fault downthrown to the south formed by
northward tilting of the forearc, or a steep southward-vergent
thrust. The Flores and Wetar Thrusts are unlikely to represent the
early stages of subduction polarity reversal, and the faults simply
represent shallow crustal features accommodating contraction in
the region between the Banda backarc and Australian continent.
Acknowledgements
We thank the consortium of oil companies who support the SE
Asia Research Group, Steve Toothill and CGGVeritas for data and for
helpful discussion, and Chris Elders, Gordon Lister and Wim
Spakman for advice and discussion. We are grateful to Mike AudleyCharles and Ron Harris for review comments that helped improve
the paper.
References
Abbott, M.J., Chamalaun, F.H., 1981. Geochronology of some Banda Arc volcanics.
Special Publication. In: Barber, A.J., Wiryosujono, S. (Eds.), The Geology and
Tectonics of Eastern Indonesia, vol. 2. Geological Research and Development
Centre, Bandung, pp. 253e268.
Abdullah, C.I., Rampnoux, J.-P., Bellon, H., Maury, R.C., Soeria-Atmadja, R., 2000. The
evolution of Sumba Island (Indonesia) revisited in the light of new data on the
geochronology and geochemistry of the magmatic rocks. Journal of Asian Earth
Sciences 18, 533e546.
Audley-Charles, M.G., 1968. The Geology of Portuguese Timor, Memoir 4. Geological
Society of London. 76p.
Audley-Charles, M.G., 1985. The Sumba enigma: is Sumba a diapiric fore-arc nappe
in process of formation. Tectonophysics 119, 435e449.
Audley-Charles, M.G., 1986a. Rates of Neogene and Quaternary tectonic movements
in the southern Banda arc based on micropalaeontology. Journal of the
Geological Society of London 143, 161e175.
Audley-Charles, M.G., 1986b. TimoreTanimbar Trough: the foreland basin to the
evolving Banda orogen. In: Allen, P.A., Homewood, P. (Eds.), Foreland Basins, vol.
8. International Association of Sedimentologists Special Publication,
pp. 91e102.
Audley-Charles, M.G., 2004. Ocean trench blocked and obliterated by Banda forearc
collision with Australian proximal continental slope. Tectonophysics 389,
65e79.
Audley-Charles, M.G., 2011. Tectonic post-collision processes in Timor. In: Hall, R.,
Cottam, M.A., Wilson, M.E.J. (Eds.), The SE Asian Gateway: History and Tectonics
of Australia-Asia Collision, vol. 355. Geological Society of London Special
Publication, pp. 241e266.
Audley-Charles, M.G., Carter, D.J., 1972. Palaeogeographical signicance of some
aspects of Palaeogene and early Neogene stratigraphy and tectonics of the
Timor Sea region. Palaeogeography, Palaeoclimatology, Palaeoecology 11,
247e264.
Audley-Charles, M.G., Carter, D.J., Barber, A.J., 1974. Stratigraphic basis for tectonic
interpretation of the Outer Banda arc, eastern Indonesia. In: Proceedings 3rd
Annual Convention. Indonesian Petroleum Association, pp. 25e44.
Audley-Charles, M.G., Harris, R.A., 1990. Allochthonous terranes of the southwest
Pacic and Indonesia. Philosophical Transactions Royal Society of London 331,
571e587.
Audley-Charles, M.G., Milsom, J.S., 1974. Comment on plate convergence, transcurrent faults, and internal deformation adjacent to Southeast Asia and the
Western Pacic by T.J. Fitch. Journal of Geophysical Research 79, 4980e4981.
Barber, A.J., Audley-Charles, M.G., 1976. The signicance of the metamorphic rocks
of Timor in the development of the Banda arc, eastern Indonesia. Tectonophysics 30, 119e128.
Bowin, C., Purdy, G.M., Johnston, C., Shor, G., Lawver, L., Hartono, H.M.S., Jezek, P.,
1980. Arc-continent collision in the Banda Sea region. American Association of
Petroleum Geologists Bulletin 64, 868e918.
Breen, N.A., Silver, E.A., Hussong, D.M., 1986. Structural styles of an accretionary
wedge south of the island of Sumba, Indonesia, revealed by SeaMARC II side
scan sonar. Geological Society of America Bulletin 97, 1250e1261.
Brown, M., Earle, M.M., 1983. Cordierite-bearing schists and gneisses from Timor,
eastern Indonesia: P-T implications of metamorphism and tectonic implications. Journal of Metamorphic Geology 1, 183e203.
Bull, S., Cartwright, J., Huuse, M., 2009. A review of kinematic indicators from masstransport complexes using 3D seismic data. Marine and Petroleum Geology 26,
1132e1151.
Burollet, P.F., Salle, C., 1981. A contribution to the geological study of Sumba
(Indonesia). In: Proceedings Indonesian Petroleum Association, 10th Annual
Convention, pp. 331e344.
Burollet, P.F., Salle, C., 1982. Histoire gologique de lile de Sumba (Indonesie).
Bulletin de la Socit Gologique de France 24, 573e580.

93

Chamalaun, F.H., Grady, A.E., von der Borch, C.C., Hartono, H.M.S., 1982. Banda arc
tectonics: the signicance of Sumba island (Indonesia). In: Watkins, J.S.,
Drake, C.L. (Eds.), Studies in Continental Margin Geology, vol. 34. American
Association of Petroleum Geologists Memoir, pp. 361e375.
Charlton, T.R., 2000. Tertiary evolution of the eastern Indonesia collision complex.
Journal of Asian Earth Sciences 18 (5), 603e631.
Charlton, T.R., 2004. The petroleum potential of inversion anticlines in the Banda
Arc. American Association of Petroleum Geologists Bulletin 88, 565e585.
Das, S., 2004. Seismicity gaps and the shape of the seismic zone in the Banda Sea
region from relocated hypocentres. Journal of Geophysical Research 109,
B12303. doi:10.1029/2004JB003192.
Dimitrov, L.I., 2002. Mud volcanoes - the most important pathway for degassing
deeply buried sediments. Earth Science Reviews 59, 49e76.
Earle, M.M., 1981. The metamorphic rocks of Boi, Timor, eastern Indonesia. Special
Publication. In: Barber, A.J., Wiryosujono, S. (Eds.), The Geology and Tectonics of
Eastern Indonesia, vol. 2. Geological Research and Development Centre, Bandung, pp. 239e251.
Effendi, A.C., Apandi, C., 1993. Geological Map of Waikabubak and Waingapu Sheets,
Nusatenggara. Geological Research and Development Centre, Bandung,
Indonesia.
Emmet, P.A., Granath, J.W., Dinkelman, M.G., 2009. Pre-Tertiary sedimentary keels
provide insights into tectonic assembly of basement terranes and present-day
petroleum systems of the East Java Sea. In: Proceedings Indonesian Petroleum Association, 33rd Annual Convention, IPA09-G-046, pp. 1e11.
Engdahl, E.R., van der Hilst, R., Buland, R., 1998. Global teleseismic earthquake
relocation with improved travel times and procedures for depth determination.
Bulletin of the Seismological Society of America 88 (3), 722e743.
Fitch, T.J., 1970. Earthquake mechanisms and island arc tectonics in the IndonesianPhilippine region. Seismological Society of America Bulletin 60, 565e591.
Fitch, T.J., 1972. Plate convergence, transcurrent faults and internal deformation
adjacent to Southeast Asia and the Western Pacic. Journal of Geophysical
Research 77, 4432e4460.
Fitch, T.J., Hamilton, W., 1974. Reply to discussion of paper by Fitch, T.J. (1972).
Journal of Geophysical Research 79, 4892e4895.
Fleury, J.-M., Pubelier, M., Urreiztieta, M., 2009. Structural expression of forearc
crust uplift due to subducting asperity. Lithos 113, 318e330.
Floodgate, G., Judd, A.G., 1992. The origins of shallow gas. Continental Shelf
Research 12, 1145e1156.
Fortuin, A.R., Roep, T.B., Sumosusastro, P.A., 1994. The Neogene sediments of East
Sumba, Indonesia - products of a lost arc? Journal of Southeast Asian Earth
Sciences 9, 67e79.
Fortuin, A.R., Roep, T.B., Sumosusastro, P.A., van Weering, T.C.E., van der Werff, W.,
1992. Slumping and sliding in Miocene and Recent developing arc basins,
onshore and offshore Sumba (Indonesia). Marine Geology 108, 345e363.
Fortuin, A.R., van der Werff, W., Wensink, H., 1997. Neogene basin history and
paleomagnetism of a rifted and inverted forearc region, on- and offshore
Sumba, Eastern Indonesia. Journal of Asian Earth Sciences 15, 61e88.
GCMT, 2011. Global CentroideMoment Tensor Catalog. www.globalcmt.org.
Gebco, 2003. IHO-UNESCO, General Bathymetric Chart of the Oceans. Digital Edition
2003, updated 2009. www.gebco.net.
Granath, J.W., Christ, J.M., Emmet, P.A., Dinkelman, M.G., 2011. Pre-Cenozoic sedimentary section and structure as reected in the JavaSPAN crustal-scale PSDM
seismic survey, and its implications regarding the basement terranes in the East
Java Sea. In: Hall, R., Cottam, M.A., Wilson, M.E.J. (Eds.), The SE Asian Gateway:
History and Tectonics of the Australia-Asia Collision, vol. 355. Geological Society
of London Special Publication, pp. 53e74.
Haile, N.S., Barber, A.J., Carter, D.J., 1979. Mesozoic cherts on crystalline schists in
Sulawesi and Timor. Journal of the Geological Society of London 136, 65e70.
Hall, R., 1996. Reconstructing Cenozoic SE Asia. In: Hall, R., Blundell, D.J. (Eds.),
Tectonic Evolution of SE Asia, vol. 106. Geological Society of London Special
Publication, pp. 153e184.
Hall, R., 2002. Cenozoic geological and plate tectonic evolution of SE Asia and the
SW Pacic: computer-based reconstructions, model and animations. Journal of
Asian Earth Sciences 20, 353e434.
Hall, R., 2009. Hydrocarbon basins in SE Asia: understanding why they are there.
Petroleum Geoscience 15, 131e146.
Hall, R., 2011. Australia-SE Asia collision: plate tectonics and crustal ow. In: Hall, R.,
Cottam, M.A., Wilson, M.E.J. (Eds.), The SE Asian Gateway: History and Tectonics
of Australia-Asia Collision, vol. 355. Geological Society of London Special
Publication, pp. 75e109.
Hall, R., Clements, B., Smyth, H.R., 2009. Sundaland: basement character, structure
and plate tectonic development. In: Proceedings Indonesian Petroleum Association, 33rd Annual Convention, IPA09-G-134, pp. 1e27.
Hall, R., Wilson, M.E.J., 2000. Neogene sutures in eastern Indonesia. Journal of Asian
Earth Sciences 18, 787e814.
Hamilton, W., 1976. Subduction in the Indonesian region. In: Proceedings, Indonesian Petroleum Association 5th Annual Convention, pp. 3e24.
Hamilton, W., 1977. Subduction in the Indonesian region. In: Talwani, M.,
Pitman, W.C. (Eds.), Island Arcs, Deep Sea Trenches and Back-Arc Basins.
American Geophysical Union, Maurice Ewing Series, pp. 15e31.
Hamilton, W., 1979. Tectonics of the Indonesian Region 345p.
Harris, R., 2003. Geodynamic patterns of ophiolites and marginal basins in the
Indonesian and New Guinea regions. In: Dilek, Y., Robinson, P.T. (Eds.), Ophiolites in Earth History, vol. 218. Geological Society of London Special Publication,
pp. 481e506.

94

J.W.D. Rigg, R. Hall / Marine and Petroleum Geology 32 (2012) 76e94

Harris, R., 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, 207e231.
Harris, R., Vorkink, M.W., Prasetyadi, C., Zobell, E., Roosmawati, N., Apthorpe, M.,
2009. Transition from subduction to arc-continent collision: geologic and
neotectonic evolution of Savu Island, Indonesia. Geosphere 5, 152e171.
Harris, R.A., 1991. Temporal distribution of strain in the active Banda orogen:
a reconciliation of rival hypotheses. Journal of Southeast Asian Earth Sciences 6,
373e386.
Harris, R.A., 1992. Peri-collisional extension and the formation of Oman-type
ophiolites in the Banda arc and Brooks Range. In: Parson, L.M., Murton, B.J.,
Browning, P. (Eds.), Ophiolites and Their Modern Oceanic Analogues, vol. 60.
Geological Society of London Special Publication, pp. 301e325.
Hendaryono, A., 1998. Etude gologique de lle de Flores. PhD thesis. Universite de
Savoie, Chambery, France, 190p.
Herrington, R.J., Scotney, P.M., Roberts, S., Boyce, A.J., Harrison, D., 2011. Temporal
association of arc-continent collision, progressive magma contamination in arc
volcanism and formation of gold-rich massive sulphide deposits on Wetar
Island (Banda arc). Gondwana Research 19, 583e593.
Hinschberger, F., Malod, J.-A., Rehault, J.-P., Villeneuve, M., Royer, J.-Y.,
Burhanuddin, S., 2005. Late Cenozoic geodynamic evolution of eastern Indonesia. Tectonophysics 404, 91e118.
Karig, D.E., Barber, A.J., Charlton, T.R., Klemperer, S., Hussong, D.M., 1987. Nature and
distribution of deformation across the Banda Arc-Australian collision zone at
Timor. Geological Society of America Bulletin 98, 18e32.
Keep, M., Longley, I., Jones, R., 2003. Sumba and its effect on Australias northwestern margin. In: Hillis, R., Mller, R.D. (Eds.), Dening Australia: The
Australian Plate as part of Planet Earth, vol. 22. Geological Society of Australia
Special Publication, pp. 309e318.
Lorenzo, J.M., OBrien, G.W., Stewart, J., Tandon, K., 1998. Inelastic yielding and
forebulge shape across a modern foreland basin: north West Shelf of Australia,
Timor Sea. Geophysical Research Letters 25, 3247.
Lunt, P., 2003. Biogeography of some Eocene larger foraminifera, and their application in distinguishing geological plates. Palaeontologia Electronica 6, 1e22.
Lytwyn, J., Rutherford, E., Burke, K., Xia, C., 2001. The geochemistry of volcanic,
plutonic and turbiditic rocks from Sumba, Indonesia. Journal of Asian Earth
Sciences 19, 481e500.
McCaffrey, R., 1988. Active tectonics of the eastern Sunda and Banda arcs. Journal of
Geophysical Research 93 (B12), 15163e15182.
McGilvery, T.A., Cook, D.L., 2003. The inuence of local gradients on accommodation space and linked depositional elements across a stepped slope prole,
offshore Brunei. In: Roberts, H.R., Rosen, N.C., Fillon, R.F., Anderson, J.B. (Eds.),
Shelf Margin Deltas and Linked Down Slope Petroleum Systems: Global
Signicance and Future Exploration Potential. GCSSEPM (Gulf Coast Section of
the SEPM), pp. 387e419.
Milsom, J., 2001. Subduction in eastern Indonesia: how many slabs? Tectonophysics
338 (2), 167e178.
Parkinson, C.D., Miyazaki, K., Wakita, K., Barber, A.J., Carswell, D.A., 1998. An overview and tectonic synthesis of the pre-Tertiary very high-pressure metamorphic and associated rocks of Java, Sulawesi and Kalimantan, Indonesia.
Island Arc 7 (1e2), 184e200.
Pirazzoli, P.A., Radtke, U., Hantoro, W.S., Jouannic, C., Hoang, C.T., Cause, C.,
Best, M.B., 1993. A one million-year-long sequence of marine terraces on Sumba
island, Indonesia. Marine Geology 109, 221e236.
Posamentier, H.W., Kolla, V., 2003. Seismic geomorphology and stratigraphy of
depositional elements in deepwater settings. Journal of Sedimentary Research
73, 367e388.
Rigg, J.W.D., Hall, R., 2011. Structural and stratigraphic evolution of the Savu basin,
Indonesia. In: Hall, R., Cottam, M.A., Wilson, M.E.J. (Eds.), The SE Asian Gateway:
History and Tectonics of Australia-Asia Collision, 355. Geological Society of
London Special Publication, pp. 225e240.
Roep, T.B., Fortuin, A.R., 1996. A submarine slide scar and channel lled with slide
blocks and megarippled Globigerina sands of possible contourite origin from the
Pliocene of Sumba, Indonesia. Sedimentary Geology 103, 145e160.
Roggeveen, P.M., 1929. Jurassic in the island of Sumba. In: Proceedings Koninklijke
Nederlandse Akademie Van Wetenschappen, pp. 512e514. Amsterdam 32.
Roosmawati, N., Harris, R., 2009. Surface uplift history of the incipient Banda arccontinent collision: geology and synorogenic foraminifera of Rote and Savu
Islands, Indonesia. Tectonophysics 479, 95e110.
Rosidi, H.M.D., Suwitodirdjo, K., Tjokrosapoetro, S., 1996. Geological Map of the
Kupang-Atambua Quadrangles, Timor. Geological Research and Development
Centre, Bandung, Indonesia.
Rutherford, E., Burke, K., Lytwyn, J., 2001. Tectonic history of Sumba island, Indonesia, since the late Cretaceous and its rapid escape into the forearc in the
Miocene. Journal of Asian Earth Sciences 19, 453e479.
Sandwell, D.T., Smith, W.H.F., 2009. Global marine gravity from retracked Geosat
and ERS-1 altimetry: ridge Segmentation versus spreading rate. Journal of
Geophysical Research 114, B01411. doi:10.1029/2008JB006008.

Satyana, A.H., Purwaningsih, M.E.M., 2011. Sumba area: Detached Sundaland


Terrane and petroleum implications. In: Proceedings Indonesian Petroleum
Association, 35th Annual Convention, IPA11-G-009, pp. 1e32.
Scotney, P.M., Roberts, S., Herrington, R.J., Boyce, A.J., Burgess, R., 2005. The development of volcanic hosted massive sulde and barite gold orebodies on Wetar
Island, Indonesia. Mineralium Deposita 40, 76e99.
Shulgin, A., Kopp, H., Mueller, C., Lueschen, E., Planert, L., Engels, M., Flueh, E.R.,
Krabbenhoeft, A., Djajadihardja, Y., 2009. SundaeBanda arc transition: incipient
continent-island arc collision (northwest Australia). Geophysical Research
Letters 36, L10304. doi:10.1029/2009GL037533.
Siebert, L., Simkin, T., 2002. Volcanoes of the World: an illustrated Catalog of Holocene
volcanoes and their Eruptions. In: Global Volcanism Program Digital Information
Series, GVP-3. Smithsonian Institution. http://www.volcano.si.edu/world/.
Silver, E.A., Reed, D., McCaffrey, R., Joyodiwiryo, Y., 1983. Back arc thrusting in the
eastern Sunda arc, Indonesia: a consequence of arc-continent collision. Journal
of Geophysical Research 88, 7429e7448.
Snyder, D.B., Prasetyo, H., Blundell, D.J., Pigram, C.J., Barber, A.J., Richardson, A.,
Tjokosaproetro, S., 1996. A dual double vergent orogen in the Banda Arc
continent-arc collision zone as observed on deep seismic reection proles.
Tectonics 15, 34e53.
Soeria-Atmadja, R., Suparka, S., Abdullah, C., Noeradi, D., Sutanto, 1998. Magmatism
in western Indonesia, the trapping of the Sumba Block and the gateways to the
east of Sundaland. Journal of Asian Earth Sciences 16, 1e12.
Spakman, W., Hall, R., 2010. Surface deformation and slab-mantle interaction
during Banda Arc subduction rollback. Nature Geoscience 3, 562e566.
doi:10.1038/ngeo917.
Standley, C.E., Harris, R.A., 2009. Tectonic evolution of forearc nappes of the active
Banda arc-continent collision: origin, age, metamorphic history and structure of
the Lolotoi Complex, East Timor. Tectonophysics 479, 66e94.
Sukamto, R., Westermann, G.E.G., 1993. Indonesia and Papua New Guinea. In:
Westermann, G.E.G. (Ed.), The Jurassic of the Circum-Pacic. Cambridge
University Press, pp. 181e193.
Tandon, K., Lorenzo, J.M., OBrien, G.W., 2000. Effective elastic thickness of the
northern Australian continental lithosphere subducting beneath the Banda
orogen (Indonesia): inelastic failure at the start of continental subduction.
Tectonophysics 329, 39e60.
Tappenbeck, D., 1940. Geologie des Mollogebirges und einiger benach-barter
Gebiete (Niederlandisch Timor). In: Brouwer, H.A. (Ed.), Geological Expedition
of the University of Amsterdam to the Lesser Sunda Islands, vol. 1. North
Holland Publishing Co., Amsterdam, pp. 1e105.
Toothill, S., Lamb, D., 2009. Hydrocarbon prospectivity of the Savu Sea basin. In:
Proceedings, Indonesian Petroleum Association 33rd Annual Convention,
IPAO9-G-013, pp. 1e12.
van Bemmelen, R.W., 1949. The Geology of Indonesia. Government Printing Ofce,
Nijhoff, The Hague. 732p.
van der Werff, W., 1995a. Cenozoic evolution of the Savu Basin, Indonesia: forearc basin
response to arc- continent collision. Marine and Petroleum Geology 12, 247e263.
van der Werff, W., 1995b. Structure and morphotectonics of the accretionary prism
along the eastern Sunda-western Banda arc. Journal of Southeast Asian Earth
Sciences 11, 309e322.
van der Werff, W., Kusnida, D., Prasetyo, H., van Weering, T.C.E., 1994. Origin of the
Sumba forearc basement. Marine and Petroleum Geology 11, 363e374.
van Weering, T.C.E., Kusnida, D., Tjokrosapoetro, S., Lubis, S., Kridoharto, P., 1989a.
Slumping, sliding and the occurrence of acoustic voids in recent and subrecent
sediments of the Savu Forearc Basin, Indonesia. Netherlands Journal of Sea
Research 24, 415e430.
van Weering, T.C.E., Kusnida, D., Tjokrosapoetro, S., Lubis, S., Kridoharto, P.,
Munadi, S., 1989b. The seismic structure of the Lombok and Savu forearc basins,
Indonesia. Netherlands Journal of Sea Research 24, 251e262.
von der Borch, C.C., Grady, A.E., Hardjoprawiro, S., Prasetyo, H., Hadiwisastra, S.,
1983. Mesozoic and late Tertiary submarine fan sequences and their tectonic
signicance, Sumba, Indonesia. Sedimentary Geology 37, 113e132.
Wakita, K., Sopaheluwakan, J., Miyazaki, K., Zulkarnain, I., Munasri, 1996. Tectonic
evolution of the Bantimala complex, south Sulawesi, Indonesia. In: Hall, R.,
Blundell, D.J. (Eds.), Tectonic Evolution of SE Asia, vol. 106. Geological Society of
London Special Publication, pp. 353e364.
Wensink, H., 1994. Paleomagnetism of rocks from Sumba: tectonic implications
since the late Cretaceous. Journal of Southeast Asian Earth Sciences 9, 51e65.
Wensink, H., 1997. Palaeomagnetic data of late Cretaceous rocks from Sumba,
Indonesia; the rotation of the Sumba continental fragment and its relation with
eastern Sundaland. Geologie en Mijnbouw 76, 57e71.
Wensink, H., van Bergen, M.J., 1995. The tectonic emplacement of Sumba in the
SundaeBanda Arc: paleomagnetic and geochemical evidence from the early
Miocene Jawila volcanics. Tectonophysics 250 (1e3), 15e30.
Wheller, G.E., Varne, R., Foden, J.D., Abbott, M.J., 1987. Geochemistry of Quaternary
volcanism in the SundaeBanda Arc, Indonesia, and three-component genesis of
island-arc basaltic magmas. Journal of Volcanology and Geothermal Research
32, 137e160.

Potrebbero piacerti anche