Vous êtes sur la page 1sur 21

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/223772058

Deformation and exhumation in Timor:


Distinct stages of a young orogeny
Article in Tectonophysics March 2010
DOI: 10.1016/j.tecto.2009.11.018

CITATIONS

READS

59

317

2 authors:
Myra Keep

David Haig

University of Western Australia

University of Western Australia

47 PUBLICATIONS 580 CITATIONS

57 PUBLICATIONS 834 CITATIONS

SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate,


letting you access and read them immediately.

SEE PROFILE

Available from: Myra Keep


Retrieved on: 08 September 2016

This article appeared in a journal published by Elsevier. The attached


copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elseviers archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright

Author's personal copy


Tectonophysics 483 (2010) 93111

Contents lists available at ScienceDirect

Tectonophysics
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / t e c t o

Deformation and exhumation in Timor: Distinct stages of a young orogeny


Myra Keep , David W. Haig
School of Earth and Environment, University of Western Australia, M004, 35 Stirling Highway, Nedlands, 6009, Australia

a r t i c l e

i n f o

Article history:
Received 13 March 2009
Received in revised form 17 November 2009
Accepted 20 November 2009
Available online 1 December 2009
Keywords:
East Timor
Biostratigraphy
Tectonics
Thrust emplacement
Early-stage orogenesis

a b s t r a c t
Timor Island, in the Outer Banda Arc, bordering the Timor Sea, preserves the orogenic product of an arc
continent collision between the Australian Plate and the Banda Arc that commenced after 10.99.8 Ma
GTS2004 but emerged above sea level only 3.1 Ma ago. The orogenic pile includes large tracts of material from
the Australian margin, including the Permian to Middle Jurassic Gondwana Megasequence and the Late Jurassic
to early Late Miocene Australian-Margin Megasequence, which occur in thrust slices. In addition, material from
the Banda Arc side of the plate margin, referred to as the Banda Terrane, occurs throughout the island and
includes both seaoor metamorphosed igneous material and cover sediments, also in thrust sheets. However
the distribution of thrust slices is unclear in many areas, perhaps because only the uppermost nappes of the
thrust pile are currently emergent and also because the thrust piles have been disrupted by later high-angle
faulting.
Evidence from East Timor suggests that the major break between deformed pre-collisional strata and the
relatively undeformed overlying deposits was during the Late Miocene (9.85.5 Ma). We present evidence for
the timing of three distinct phases of orogenic development, as determined from East Timor, including initial
collision and emplacement of the early nappes creating loading and diapirism (within the 9.85.5 Ma interval),
a tectonic quiet interval (5.5 Ma4.5 Ma) that extended for about a million years during the middle of the
collision and may represent the time of locking of the subduction system, and a post 4.5 Ma phase of uplift,
unroong and further diapirism in response to isostatic rebound. Our conclusions offer an alternative model for
the evolution of this part of the Banda Arc.
2009 Elsevier B.V. All rights reserved.

1. Introduction
The island of Timor, in the Outer Banda Arc (Figs. 1 and 2), is
perhaps the best place on earth to study young arccontinent collision.
The distance between the northern edge of the continent and the arc at
their closest point (from Dili to Ata'uro, in East Timor) is less than
25 km (Fig. 2). Between them a steep gravity anomaly (Chamalaun
et al., 1976) coincides with the 3 km-deep trench that occupies the
narrow strait, and has been postulated to represent a steep fault
separating the continental and oceanic crust (Chamalaun et al., 1976),
or possibly even the northern edge of the Australian Plate (AudleyCharles, 2004).
The orogenic product of the collision between the Australian Plate
and Banda Arc, a mountain belt larger in extent than the Swiss Alps
(Audley-Charles, 1981; Brunnschweiler, 1978), began emerging from
the sea only 3 Ma ago (Haig and McCartain, 2007). This young age
means that the rocks exposed on Timor Island likely represent only
the higher structural levels of a fold-and-thrust belt that is still largely
not exposed (Brunnschweiler, 1978). Regionally pervasive vegetation,
and the lack of any deep excision of the orogen, mean that geological

Corresponding author.
E-mail address: myra.keep@uwa.edu.au (M. Keep).
0040-1951/$ see front matter 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.tecto.2009.11.018

contacts are not always well-exposed, and this has caused a history of
contrasting interpretations, and some controversy, since structural/
tectonic geological research was rst conducted in the area (Hirschi,
1907). In addition, since geologic expeditions began, the island has
been divided into two geo-political entities, namely Dutch Timor
(Indonesian West Timor) and Portuguese Timor (East Timor) (Fig. 2),
which has tended to focus research efforts on either West or East
Timor; rarely have investigations crossed the boundary. This has
imparted some bias into published interpretations, which was greatly
enhanced by the inability to access East Timor for geological research
for most of the period from 1975 to 2002. Signicantly more research
has been conducted in West Timor (see overview below). In addition,
most of the research expeditions tended to focus on small areas, rather
than the entire orogen, for reasons of access, topography, logistics and
funding. The natural tendency to extrapolate small-scale ndings into
regional interpretations, and then to extrapolate interpretations from
West Timor into East Timor, has fed controversy regarding geological
interpretations.
Over the last few years, we have made considerable revisions of
the Cenozoic stratigraphy in East Timor based on detailed biostratigraphy (Haig and McCartain, 2007; Haig et al., 2007; Haig et al., 2008).
In this paper we place the new stratigraphic ndings into a tectonic
framework. This allows us to re-examine the various tectonic models
proposed for Timor Island, which in turn may have implications for

Author's personal copy


94

M. Keep, D.W. Haig / Tectonophysics 483 (2010) 93111

Fig. 1. Outline map of the Australian continental margin, modied from Heap and Harris (2008). Location of offshore plateaux are shown, as well as the collided segment of the
margin in the Timor region.

Fig. 2. Location map of Timor Island, showing the distribution of the main geological megasequences, and the location of specic towns and geological formations discussed in the
text. The location of the probable northern boundary of the Australian Plate is also shown, and is geographically known as the Wetar Strait. This corresponds to the geophysical
anomaly of Chamalaun et al. (1976). The inset map shows the tectonic setting of Timor Island in the Outer Banda Arc. Inset shows the regional physiographic elements. 1 = Timor
Sea, 2 = Browse Basin.

Author's personal copy


M. Keep, D.W. Haig / Tectonophysics 483 (2010) 93111

much of the greater regional geology. Many of the smaller islands in


the region were mapped using the same tectonostratigraphic units
observed in East Timor, and these also may be in need of signicant
stratigraphic revision.

2. Overview of previous work


Given the highly contentious nature of geological interpretations
of Timor Island, it is pertinent to present a short history of research
on Timor, highlighting especially the previous work in East Timor
compared to West Timor. To avoid the use of conicting terminology,
we use the present-day names East Timor and West Timor to refer
to work conducted in Portuguese and Dutch/Indonesian Timor respectively. It should be noted that for West Timor in particular this
summary is not intended as comprehensive, as geological work has
been conducted there relatively continuously since the late 1800s.
Early geological investigations on Timor Island focussed on
palaeontology (e.g. Beyrich, 1865; Rothpletz, 1891, 1892). For West
Timor, Alpinic overthrust tectonics were rst proposed by Wanner
(1913), and other workers (Molengraaf, 1912, 1915; Krumbeck, 1921;
Brouwer, 1942; Simons, 1940; Tappenbeck, 1940; De Roever, 1940;
De Bruyne, 1941; Van West, 1941; De Waard, 1954a, b, c, d; 1955a, b;
1956; 1957). In 1967, International Oils commenced exploration in
West Timor, which resulted in two theses on West Timor (Giani,
1971; Kenyon, 1974). During the post-1975 period, research continued in West Timor (e.g. Sartono, 1975; Rosidi et al., 1979; Hamilton,
1979; Browne and Earle, 1983; Kristan-Tollman, et al., 1987;
Sopaheluwakan et al., 1989; Wensink and Hartosukohardjo, 1990;
Bird and Cook, 1991; Sawyer et al., 1993; Sani et al., 1995; Sashida
et al., 1996, 1999; Martini et al., 2000; Harris and Long, 2000), and
included a number of theses from the University of London (e.g.
unpublished theses by Cook, 1986; Bird, 1987; Charlton, 1987; Harris,
1989; Barkham, 1993; Rose, 1994). The concentration of research
efforts in West Timor during this period, and the corresponding
dearth of data for East Timor, has led to speculative correlations of
geological structures and units from West Timor into East Timor, (e.g.
Harris et al., 1998; Charlton, 2002; Villeneuve et al., 2005).
For East Timor, Hirschi (1907) provided one of the earliest
descriptions of deformation, and referred to possible klippes, implying
thrust deformation. Subsequent work for resource exploration (Wittouck, 1937; Grunau, 1953, 1956, 1957a, b) was mainly reconnaissance
in nature, but included a gravity survey (Teixeira, 1952). Van Bemmelen
(1949) summarized the history of the early geological investigations in
the Outer Banda Arc, including Timor, and provided a more modern
account of the stratigraphy and structure of the region up to that
time. Further geological reconnaissance work in East Timor in 1955
(Gageonnet and Lemoine, 1957a,b,c, 1958), preceded petroleum
exploration by Timor Oil in 1956 (Brunnschweiler, 1978). From mid1959 Audley-Charles conducted 28 months of eld mapping and subsequent analysis for Timor Oil, which culminated in his PhD thesis and
subsequent 1968 memoir (Audley-Charles, 1968).
Further eldwork by Audley-Charles and associates in JulyAugust
1973 culminated in papers that included changes to Audley-Charles'
(1968) published map (e.g. Audley-Charles et al., 1974; Carter et al.,
1976; Barber and Audley-Charles, 1976; Barber et al., 1977).
In 1974, Woodside-Burmah Oil, N. L., began to explore concessions
in both West and East Timor. During this work, Crostella and Powell
(1975) noted that it was almost impossible to dene normal lithostratigraphic units in the conventional sense, due to the fact that
many previously mapped stratigraphic contacts were in fact tectonic.
Independently during the early 1970s, the Flinders University of
Australia undertook a re-examination of some of the Audley-Charles
(1968) thrust boundaries (Grady, 1975; Grady and Berry, 1977) and
palaeomagnetic (Chamalaun, 1977) and gravity studies (Chamalaun
et al., 1976), as well as a detailed examination of the Aileu Formation

95

(e.g. Berry, 1981; Berry and Grady, 1981; Berry and Jenner, 1982;
Berry and McDougall, 1986).
From late 1975, the political situation in East Timor prevented
almost any further eld geological investigations other than those
conducted by Indonesian geologists (e.g. Bachri and Situmorang,
1994), although some access to geologists was granted in the 1990s
(e.g. Hunter, 1993; Reed et al., 1996; Harris et al., 2000).
Despite the lack of eld work in East Timor, regional work
elsewhere in the Banda Arc continued, with offshore data acquisition
(e.g. Bowin et al., 1980; Silver et al., 1985; Woodside et al., 1989;
Snyder et al., 1996; Honthaas et al., 1998) and tectonic modelling and
plate reconstruction (e.g. Barber and Audley-Charles, 1976; Barber
et al., 1977; Chamalaun and Grady, 1978; Audley-Charles et al., 1979;
Earle, 1979; Audley-Charles, 1981, 1983, 1985; McCaffrey et al., 1985;
Audley-Charles, 1986a,b; Barber et al., 1986; Karig et al., 1987; Price
and Audley-Charles, 1987; Audley-Charles et al., 1988; Charlton,
1989; Rangin et al., 1990; Audley-Charles and Harris, 1990; Daly et al.,
1991; Hall, 1997; Harris et al., 1998; Hall, 2002; Audley-Charles, 2004;
Audley-Charles, 2004).
The rst of the new era publications on East Timor (Monteiro,
2003; Roniewicz et al., 2005) concentrated on the Triassic of East Timor.
Since 2003, a number of theses and papers have been produced by
Australian and other researchers (e.g. Keep et al., 2005; Villeneuve et al.,
2005; Falloon et al., 2006; McCartain et al., 2006; Harris, 2006; Haig and
McCartain, 2007; Haig et al., 2007; Lisboa et al., 2007; Haig et al., 2008;
Standley and Harris, 2009; Nugroho et al., 2009; Roosmawati and Harris,
2009; Keep et al. 2009).
3. Tectonic overview
Geologically Timor Island displays juxtaposed rocks from both the
Australian and Banda Arc sides of the plate boundary (e.g. AudleyCharles, 1968). Australian continental crust extends at least to the north
coast of Timor (e.g. Chamalaun et al., 1976), and the current limit of
the Australian Plate likely coincides with the steep gravity low and
signicant topographic low (N3 km-deep Wetar Strait) that lies north of
the northern Timor coast (Audley-Charles, 2004) (Fig. 2). Timor does
not display structurally or stratigraphically coherent, strike parallel
divisions as are common in more mature young orogens, such as Taiwan
(e.g. Huang et al., 2000; Chang et al., 2003). Instead there is a broad
decrease in metamorphic grade from north to south and a tectonostratigraphy that is mostly separated into Banda Terrane rocks and
megasequences of the AustralianGondwanan continent (Fig. 2).
In general, most workers (e.g. Hirschi, 1907; Wittouck, 1937;
Grunau, 1953; Gageonnet and Lemoine, 1958; Marks, 1961; AudleyCharles, 1968; Carter et al., 1976; Barber et al., 1977; Brunnschweiler,
1978; Audley-Charles, 1981; Sopuhaluwekan et al., 1989; Harris,
1991; Charlton, 1991; Sani et al., 1995; Villeneuve et al., 1999; Keep
et al., 2005) agree on the presence of crustal shortening and thrust
nappes in Timor (see Grady, 1975 and Crostella and Powell, 1975
for opposing views), but disagree, sometimes signicantly, on the
exact nature of contacts (Grady, 1975; Grady and Berry, 1977), the
origin (Australian or Banda Arc) of some apparent thrust slices (e.g.
Charlton, 2002), the age of units (e.g. Haig et al., 2008), the timing of
deformation (e.g. Audley-Charles, 2004), and the timing of collision
(e.g. compare Hall, 2002; Audley-Charles, 2004 and Roosmawati and
Harris, 2009 with Reed, 1985; Berry and McDougall, 1986; Richardson
and Blundell, 1996; Rutherford et al., 2001; Keep et al., 2002, 2003;
Haig and McCartain, 2007). This is compounded by the lack of exposure of many presumed thrust contacts (e.g. Grady and Berry,
1977), and the relative simplicity of many of the models (e.g. AudleyCharles, 1968; Tobing, 1989; Harris, 1991; Harris et al., 1998). Some
models propose a coherent series of nappes (e.g. Audley-Charles,
1968; Barber et al., 1977; Brunnschweiler, 1978; Tobing, 1989) or
alternatively propose a stylised series of thrust slices (Harris, 1991;
Harris et al., 1998), both of which are hard to reconcile on the ground.

Author's personal copy


96

M. Keep, D.W. Haig / Tectonophysics 483 (2010) 93111

In particular the role of potentially overturned stratigraphy in thrust


nappes seems to have been little considered (Keep et al., 2009).
We present a series of stages in the evolution of the eastern part of
Timor Island (East Timor), as dened by the stratigraphy, palaeogeography, and deformation. We use the stratigraphic nomenclature of
megasequences (Fig. 3) that was outlined by Haig et al. (2007), and
the Geological Time Scale (which we refer to as GTS2004) developed
in Gradstein et al. (2004).
4. Pre-collisional palaeogeography
The pre-collisional palaeogeography of the Australian margin that
included the region of Timor (Chamalaun et al., 1976; Chamalaun,
1977) was established through the breakup of Gondwana during the
Late Jurassic (Pigram and Panggabean, 1984; Mller et al., 1998; Li and
Powell, 2001; Longley et al., 2002). The age of breakup in the Timor
region has been estimated at 155 Ma (Mller et al., 1998). Prior to
breakup, the Timor region formed part of a broad system of narrow
interior-rift basins and broader interior-sag basins that extended
south from the north-eastern margin of Gondwana for several
thousand kilometres reaching deep into the Gondwanan interior
(Audley-Charles, 1983; Audley-Charles et al., 1988; Metcalfe, 1996;
Harroweld et al., 2005; Metcalfe, 2006).
The modern day geometry of the Australia rifted margin comprises
a series of salients and embayments (Fig. 4), proposed to correspond
to the rifted upper- and lower-plate margins (cf. Lister et al., 1986) of
the northern Australian margin (Harroweld and Keep, 2005). In the
region of Timor the upper-plate salient has now collided with the
Banda Arc (Harroweld and Keep, 2005; Haig and McCartain, 2007),
causing tectonic modications in this region, whereas the present-day
Exmouth Plateau upper-plate salient preserves much of the original
rift-margin architecture (Fig. 4). The irregular shape of the ocean
continent boundary on the Exmouth Plateau is clearly observable
on bathymetric and gravity images of the region (Fig. 4), and can be
used as an analogue for the geometry of the now-collided Timor Sea
segment (e.g. McCaffrey et al., 1985). Note from Fig. 4 that the
present-day Exmouth Plateau upper-plate salient extends more than
500 km further out than the adjoining lower-plate embayment
represented by the Browse Basin (Harroweld and Keep, 2005), and
similarly the Wallaby Plateau protrudes almost 900 km further out
than the easterly-adjacent continental crust. Using the shape of the

Fig. 3. Generalised stratigraphic sequences of Timor. Dark grey boxes correspond to the
para-autochthon of Audley-Charles (1968), whilst the light grey boxes and white boxes
correspond to his autochthon and allochthon respectively. Modied from Haig et al.
(2007).

present-day oceancontinent boundary on the Exmouth Plateau


upper-plate (Fig. 1) as an analogue for the now-collided Timor sea
upper-plate (Fig. 4), the geometry of the margin prior to collision
would be highly irregular (even more so if the geometry of the
Wallaby Plateau rather than the Exmouth Plateau is used), with
outlying plateaux extending up to 900 km further north than adjacent
areas of continental crust.
These submarine plateaux (Fig. 1) are underlain by attenuated
continental crust (Fig. 5) contiguous with the thicker crust of the
mainland, and subsided to mid-bathyal water depths during the Early
Cretaceous (Exon et al., 1992; Heap and Harris, 2008). Heap and
Harris (2008) showed that the plateaux facing the Indian Ocean have
a trough developed on their eastern side adjacent to the continental
shelf (Fig. 5a), whereas on their oceanic sides continental crust gives
way to volcanic-margin crust and then to oceanic crust (Fig. 5b)
(Direen et al., 2008; Rey et al., 2008). The notion that the Timor region
was originally a plateau similar to the Exmouth Plateau (Haig and
McCartain, 2007), on the basis of both likely margin geometry and the
reconstructed sedimentary succession, contrasts with an alternate
scenario of Audley-Charles (1988, 2004), that the collided crust was a
relatively narrow continental slope and rise.
The Exmouth Plateau analogue rather than a narrower slope and
rise scenario is supported by six lines of argument:
1. The Gondwana and Australian-Margin megasequences in Timor
(Haig and McCartain 2007; Haig et al., 2007) broadly correspond
in facies and age to equivalent successions on ExmouthWombat
Plateau (summarized in Exon et al., 1992). The relative thicknesses of
these units on Exmouth Plateau (Fig. 5b) seem consistent with the
relative thicknesses interpreted for Timor (Haig and McCartain 2007).
2. At several localities in East Timor, the deformed Australian-Margin
Megasequence is closely associated with outcrops of shales containing a Malayamaorica-Belemnopsis molluscan fauna that is known
elsewhere only from shelf deposits that accumulated on the
AustralianAntarcticZealandia margin (Haig and McCartain 2007).
3. The Australian-Margin Megasequence in Timor consists of Late
Jurassic shelf facies overlain by deep-water pelagic deposits of the
Cretaceous to early Late Miocene (Haig and McCartain, 2007),
suggesting subsidence during the earliest Cretaceous comparable
to that evidenced on Exmouth Plateau.
4. All of the known East Timor localities of the Australian-Margin
Megasequence, both on the north as well as the south side of the
island, are closely associated with outcrops of the upper part of
the Gondwana Megasequence (Haig and McCartain, 2007). This
suggests that the Australian-Margin Megasequence originally
overlay the Gondwana Megasequence throughout the region.
5. Clear evidence for turbidites, as it would occur on a continental
rise and adjacent abyssal plain (e.g. as present in the Cretaceous
Cenozoic succession on the adjacent Argo Abyssal Plain; Gradstein
et al., 1990) has not been documented in the carbonate pelagite
succession of the Australian-Margin Megasequence in East Timor
(Haig and McCartain, 2007). Chaotic folding and mixing of microfossils in these deposits can be explained by diapirism within uidsaturated soft sediment during early stages of orogenesis rather
than by slumping and reworking on a slope.
6. Based on a seismic interpretation presented by Petkovic et al. (2000),
the collective thickness of the Gondwana and Australian-Margin
megasequences remains constant across the southern ank of the
Timor Trough but becomes obscured at the orogenic deformation
front close to the axis of the Trough. Compared to its thickness on the
Australian Shelf, the interpreted Lower Cretaceous part of the
Australian-Margin Megasequence appears to thin substantially on
the upper part of the southern slope of the trough.
We therefore suggest that the pre-collisional paleogeography
resembled the reconstructed rift margin shown in Fig. 4, based on

Author's personal copy


M. Keep, D.W. Haig / Tectonophysics 483 (2010) 93111

97

Fig. 4. Possible geometries of the segmented Australian margin prior to collision. Solid portion of the line indicates present-day geometry of the Carnarvon Basin. a. Segment shown
as grey (present-day geometry) has been used as a template for the pre-collisional Timor Plateau, shown as dashed lines. b. Entire geometries of the Wallaby and Exmouth plateaux
(present-day) have been used as a template for the pre-collisional Timor Sea, shown as dashed lines. Note in both cases signicant salients and embayments occur along the entire
collided margin. c. Marine topography (after Sandwell and Smith, 1997). Note the prominent salients of the Exmouth and Wallaby plateaux, and the position of Sumba and it's high
topography in relation to the present-day Wallaby plateau. d. Free air gravity derived from satellite altimetry (modied from Sandwell and Smith, 1997). Red areas are gravity highs
and purple areas are gravity lows. Note the prominent gravity high underneath Sumba Island, a possible remnant of underplated Australian continental crust from an outlying
Wallaby-type plateau.

sedimentological evidence presented above, and using the presentday non-collided geometries of the Australian margin as an analogue.
5. Phases of collision
Although most investigators agree that initial northward subduction
of the Australian continental margin beneath the Banda Arc leads to
jamming of subduction, overthrusting and uplift of Timor, the nature
and timing of key events are in dispute. Table 1 shows a comparison
between events based on East Timor stratigraphy (Haig and McCartain,
2007) and those outlined in Audley-Charles (2004). The ages presented
by Haig and McCartain (2007) have been slightly modied to take into
account the planktonic foraminiferal datum levels (Fig. 6) assessed by
Sinha and Singh (2008). The chronometry outlined by Audley-Charles
(2004) had its legacy several decades ago in planktonic foraminiferal
zonal determinations made by D. J. Carter, for which the documentation
is largely unpublished. Many of the age determinations were made
on successions in West Timor. Because of the importance of these
age determinations in the development of prevailing views of Timor
orogenesis, a summary of available biostratigraphic determinations
from West Timor that are either published or presented in the unpublished PhD thesis of Kenyon (1974) is presented in Table 2 and
discussed in relevant sections below.
5.1. Initial collision
We place the timing of initial collision after the age of the youngest
unit of the deformed Australian-Margin Megasequence that we have
found in East Timor (Fig. 3). This unit is a pelagite, with known

outcrop at several widely separated localities, belonging to planktonic


foraminiferal Zone N15 (Figs. 6 and 7) representing the lowest part of
the Upper Miocene (Haig and McCartain, 2007), implying collision
after 10 Ma.
In West Timor, deformed strata as young as Early Pliocene have
been reported from the Kolbano region (Fig. 7). Audley-Charles (2004)
regarded N18 pelagic lutites [latest Miocene according to GST2004]
in the Kolbano succession of West Timor as the youngest deformed
parautochonous rocks. These limestones were described by Kenyon
(1974) as the Sabaoe Limestone Formation of the Kolbano region
from which D. J. Carter had recorded a general N18 to N21 age range
[latest Miocene to Late Pliocene; GTS2004], although Kenyon regarded
them apparently as N21 [Middle to Late Pliocene; GTS2004] because of
a lithological resemblance to the Sabaoe Limestone in the Central Basin
of Timor. Charlton and Wall (1994) renamed the unit of strongly
deformed limestones in the Kolbano area as the Siu Formation and
cited samples belonging to zones N18N19/20 (latest Miocene to early
Late Pliocene according to GTS2004; Table 2, rows 1 to 3) as well as
one sample within an imprecise Middle Miocene to Lower Pliocene
range and one sample with a broad Lower Miocene range based on
palynomorphs. These ranges led Charlton and Wall (1994, Fig. 3;
followed by Charlton, 2002) to conclude that the Siu Formation ranges
throughout the Miocene and into the lowest Pliocene. They noted that
Audley-Charles and Carter (1972) had described it as the Bolan facies.
Charlton and Wall (1994) also suggested a possible correlation
between the Siu Formation and the Aliambata Limestone described
from East Timor by Audley-Charles (1968).
The youngest age we have determined for the Aliambata
Limestone in East Timor is earliest Late Miocene (Zone N15; Haig

Author's personal copy


98

M. Keep, D.W. Haig / Tectonophysics 483 (2010) 93111

Fig. 5. a. Geomorphological map of Exmouth Plateau (see Fig. 1 for location), considered an analogue for a pre-collisional Timor plateau. The geomorphic mapping follows Heap and
Harris (2008). The line XX indicates the approximate position of the section show in b. b. A schematic cross-section of Exmouth Plateau following Direen et al. (2008) and Petkovic
et al. (2000). The stratigraphy on Exmouth Plateau seems to be analogous to the pre-collisional stratigraphy in East Timor.

and McCartain, 2007), representing what we think is the youngest


part of the Australian-Margin Megasequence. The youngest age of the
Bolan facies determined by Audley-Charles and Carter (1972) and the
species listed by them (Table 2, row 4) are consistent with the age and
fauna of the Zone N15 deposits in East Timor (taking into account
recent changes in time scale). In the published biostratigraphic record,
there seems to be no denite indication that the Upper Miocene
interval included in planktonic foraminiferal zones N16 and N17 is

present in the Siu Formation of the Kolbano region. A more detailed


biostratigraphic study of the Siu Formation is not available to conrm
the full stratigraphic range of the formation and the presence of any
major unconformities.
Taking 9.8 Ma, our youngest age for the deformed Australian Margin
Megasequence, as the oldest limit for major collisional deformation is
consistent with several other datasets. Berry and McDougall (1986)
proposed a collisional age of 8 Ma based on 40Ar/39Ar and K/Ar ages for

Table 1
Comparison of tectonic events recognized in the present study with those outlined by Audley-Charles (2004).
Present study slightly modied from Haig and McCartain (2007)
which updated Pliocene chronometry based on Fig. 6

Audley-Charles (2004)

After 9.8 Ma: Initial collision of Timor plateau (Australian continental terrace/plateau)
with Banda Arc
9.85.5 Ma: Phase of crustal shortening and loading

12 Ma: Initiation of Banda Arc (based on oldest volcanics known in Arc)

5.54.5 Ma: Tectonic quiet interval due to locking of the subduction system
4.53.1 Ma: Uplift resulting in initial emergence of Timor island due to subducting
slab detachment and isostatic rebound
Post 3.1 Ma: Additional uplift and late extension

123.5 Ma: Subduction of northward-dipping Australian continental margin under


the Banda Arc accretionary prism formed by Australian continental rise sediments
52 Ma: Australian continental slope entered Banda subduction trench
3.52 Ma: Australian continental margin and volcanic arc collision took place and
subduction was locked
3.52 Ma: Uplift of Timor was initiated

Author's personal copy


M. Keep, D.W. Haig / Tectonophysics 483 (2010) 93111

99

Fig. 6. Planktonic foraminiferal datum levels used for age determination of Pliocene strata in Timor. Zone denitions follow Blow (1969) except for the following revisions (for zones
designated by r after zonal number): N19rN20r boundary dened at the lowest occurrence of Globorotalia crassaformis; N22rN23r boundary dened at the highest occurrence of
Globorotalia tosaensis. Species diagnoses follow Kennett and Srinivasan (1983). Blow (1969) used the rst appearance of Globorotalia pseudopima to dene the N19N20 boundary,
but this datum has been shown as unreliable (Brnnimann and Resig, 1971).

hornblendes from the Alieu Formation, indicating retrograde metamorphism in this part of the Gondwana Megasequence at this time (see
Figs. 2 and 3). An approximately 8 Ma age for collision has also been
proposed on the basis of mass balance sediment calculations (Richardson and Blundell, 1996), and the age of deformation in the Timor Sea
(Keep et al., 2002; Harroweld et al., 2003). It also coincides with the
initiation of uplift of Sumba Island from 7 Ma (van der Werff et al., 1994;
Fortuin et al., 1994, 1997; Rutherford et al., 2001). Harris et al. (2000)
also recognise a 98 Ma event, which they attribute to tectonic burial
beneath the Banda Terrane nappe. A Late Miocene age for collision also
is consistent with inversion of structures seen in basins along the
western margin of Australia as far south as the Southern Carnarvon
Basin (Hocking et al., 1987; Keep et al., 2002; Longley et al., 2002).
The initial collision between 9.8 Ma and 5.5 Ma appears to mark
the onset of crustal shortening and loading in East Timor, representing
the start of the orogenic process.

5.2. 9.8 to 5.5 Ma non-deposition/crustal shortening


The interval 9.8 Ma to 5.5 Ma is marked by a lack of deposition in
East Timor, between the youngest unit of the deformed AustralianMargin Megasequence (Zone N15; earliest Late Miocene) and the
youngest beds of the Synorogenic Megasequence (latest Miocene
Zone 18 in the basal beds of the type section of the Viqueque
Formation in East Timor; Haig and McCartain, 2007). Kenyon (1974)
also noted the absence of the intervening zones N16 and N17 in West
Timor, although he ascribed this to tectonic shearing. In contrast
to the pre-collision limestones of zone N15 that are indurated,
stylolitized, and highly deformed, the basal beds of Viqueque
Formation (N18N19) are friable gently deformed chalk.
In terms of similar tectonostratigraphic attribution, AudleyCharles (1986a) recognized an unconformity between the youngest
unit of the deformed parautochthon (the deformed N18 lutites of

Author's personal copy


100

M. Keep, D.W. Haig / Tectonophysics 483 (2010) 93111

Table 2
Summary of zonal determinations for units critical for understanding the Pliocene tectonic history of West Timor.
Stratigraphic unit

Planktonic foraminifera determined by Simon Robertson


Indonesia (13) and D.J. Carter (413). Identications as
in original lists

Original zonal assignation

1. Siu Formation locality TIM 49


(Charlton and Wall, 1994)

Globigerinoides trilobus, G. immaturus, G. extremus,


G. obliquus, G. ruber, Globoquadrina altispira
Neogloboquadrina acostaensis, Globorotalia menardii,
G. miocenica, G. margaritae, G. tumida

2. Siu Formation locality SIU 18


(Charlton and Wall, 1994)

Orbulina universa, Sphaeroidinella subdehiscens, Globorotalia


menardii, G. acostaensis, G. miocenica, G. tumida,
Globigerinoides immaturus, G. extremus, Globoquadrina
altispira, Globigerina seminulina
Globigerina trilobus, G. immaturus, Globoquadrina altispira,
Globorotalia tumida

Charlton and Wall (1994): Assemblage does not include boundary


N18N19
marker species for zones above N18.
The last appearance of G. margaritae
lies within zone N20r (Fig. 6); the last
appearance of G. altispira lies at the base
of zone N21 (Fig. 6).
Charlton and Wall (1994): Assemblage does not include boundary
N18N19/20
marker species for zones above N18.
The last appearance of G. altispira lies
at the base of zone N21 (Fig. 6).
Charlton and Wall (1994): Assemblage does not include boundary
N18N19/20
marker species for zones above N18.
The last appearance of G. altispira lies
at the base of zone N21 (Fig. 6).
Audley-Charles and Carter Species are consistent with the range
(1972): Lower Miocene
noted by Audley-Charles and Carter,
extending into the Middle (1972) (see discussion in text)
Miocene

3. Siu Formation locality SIU 33


(Charlton and Wall, 1994)

4. Bolan facies, River Bolan near Kolbano Globigerina praebulloides pseudociperoensis, Globigerinoides
(Audley-Charles and Carter, 1972)
quadrilobatus primordius, Globoquadrina altispira altispira,
G. altispira globosa, G. dehiscens advena, G. dehiscens
dehiscens, G. venezuelana s.l., Globorotalia cultrata
cultrata, Orbulina universa, Sphaeroidinellopsis seminulina
seminulina, S. seminulina kochi.
5. Type section of Batu Patih Limestone
Globorotalia cultrata menardii, G. tumida tumida (sinistral),
(Kenyon, 1974)
G. subcretacea, G. acostaensis pseudopima, G. cultrata cultrata,
Globigerina nepenthes, Pulleniatina spp., Sphaeroidinella
dehiscens, Globigerinoides gomitulus, G. ruber, G. conglobatus,
G. quadrilobatus, G. obliquus obliquus.
6. Lower part of Batu Putih Limestone Globorotalia cultrata menardii, G. tumida tumida (sinistral),
G. cultrata cultrata, Globigerina nepenthes, Pulleniatina
(Audley-Charles, 1986a)
primalis, Globigerinoides gomitulus, G. ruber, G. conglobatus,
G. quadrilobatus sp., G. obliquus obliquus, Sphaeroidinellopsis
subdehiscens subdehiscens, Sphaeroidinellopsis seminulina
kocki and Sphaeroidinellopsis subdehiscens paenedehiscens.
7. Fatu Laob Member of Batu Putih
Globoquadrina dehiscens dehiscens, Sphaeroidinellopsis
Limestone (Audley-Charles, 1986a)
seminula seminula, Globigerina decoraperta, Pulleniatina
obliquiloculata praecursor
8. Sabaoe Limestone Formation
(Kenyon 1974)

9. Oemenu Sandstone Member


(Kenyon 1974)

10. Nuanain Sandstone Member


(Kenyon 1974)

11. Tanah Putih Marl of Noele


Marl Formation (Kenyon 1974)

Orbulina universa, Globorotalia cultrata menardii,


G. subcretacea, G. acostaensis pseudopima, G. borealis,
G. cultrata cultrata, Pulleniatina spp., Sphaeroidinella
dehiscens, Globigerinoides gomitulus, G. conglobatus gp.,
G. quadrilobatus gp.
Orbulina universa, Globorotalia cultrata menardii, G. tumida
tumida (sinistral). G. subcretacea, G. borealis, G. acostaensis
pseudopima, Pulleniatina spp., Sphaeroidinella dehiscens,
Globigerinoides gomitulus, G. ruber, G. quadrilobatus gp.,
G. conglobatus gp., Globigerina bulloides, G. parabulloides,
G. rubescens, G. juvensis, G. calida praecalida.
Orbulina universa, Globorotalia cultrata menardii,
G. subcretacea, G. acostaensis pseudopima, G. borealis gp.,
G. tumida tumida (sinistral), G. tosaensis tenuitheka,
Pulleniatina spp., Globigerinoides gomitulus, G. ruber,
G. conglobatus gp., G. quadrilobatus gp., Globigerina bulloides,
G. parabulloides, G. rubescens, G. juvensis, G. calida praecalida.
Globigerina bradyi, G. juvensis, G. bulloides, G. cf. bulloides,
Globigerinoides quadrilobatus gp., G. gomitulus, G. ruber gp.,
Globorotaloides variabilis hexagona, Globigerinita
ambitacrena, Globorotalia scitula scitula, S. subcretacea,
G. cultrata exilis, G. crassaformis crassaformis

the Kolbano region, discussed above) and the oldest unit of the
autochthon (i.e. the oldest synorogenic unit in our classication)
which he considered to belong to zone N21 (Sabaoe Limestone or Fatu
Loab Member of the Batu Putih Formation; Late Pliocene according to
GTS2004). Based on planktonic foraminifera (see Table 2, rows 7 and
8), he considered that zone N20 (late Early to early Middle Pliocene
according to GTS2004) was missing within the Batu Putih Formation,
and placed the lower part of the Batu Putih Formation (zones N18
N19; latest Miocene to Early Pliocene according to GTS2004) as the
youngest member of the allochthon (i.e. the Banda Terrane in our
classication). The Miomaffu Tuff of the Noiltoko area of West Timor
(Fig. 7), originally described by Van West (1941) and attributed by

Comments on zonal determination

Kenyon (1974), p. 27:


N18lowermost N21

Assemblage does not include boundary


marker species for zones above N19r.
The last appearance of G. nepenthes lies
within zone N20r (Fig. 6)

Audley-Charles (1986a):
N18N19

Assemblage does not include boundary


marker species for zones above N18.
The last appearances of G. nepenthes
and Sphaeroidinellopsis spp. lie within
Zone N20r (Fig. 6).

Audley-Charles (1986a):
N21

Assemblage does not include boundary


marker species. The last appearance of
Sphaeroidinellopsis spp. lies within
uppermost part of Zone N20r (Fig. 6).
Assemblage does not include boundary
marker species for zones above N19r
nor any of the species with datum
levels plotted in Fig. 6.

Kenyon (1974), p. 52:


lower part of N21

Kenyon (1974), p. 78:


young N21

Assemblage does not include boundary


marker species for zones above N19r
nor any of the species with datum l
evels higher than the basal N19r
boundary plotted in Fig. 6.

Kenyon (1974), p. 92:


N22

Assemblage does not include the lower


boundary marker species for zone N22.
Other species are consistent with a
position level somewhere between
the base of N18 and the top of N22.

Kenyon (1974), p. 128:


N22

Assemblage does not include boundary


marker species for zones above N20r
nor any of the species with datum
levels higher than the basal N20r
boundary plotted in Fig. 6

Audley-Charles (1986a) to zone N17 (Late Miocene), was also


considered part of the allochthon. This view developed from the
work of Carter et al. (1976) and was echoed in Audley-Charles (2004)
with one important difference. The lower part of the Batu Putih
Limestone was now described as the oldest autochthonous rocks
resting unconformably on the deformed rocks and was placed in
zones N1819 which is between 5.0 and 3.5 Ma (Audley-Charles,
2004, p. 73). According to the datum levels plotted against GTS2004 in
Fig. 6, zone N1819 is now dated between 5.5 and 4.5 Ma.
Species lists (Table 2, rows 513) that Audley-Charles (2004) used
for age determination of critical units do not contain denitive
evidence for a major unconformity in the Pliocene succession. The

Author's personal copy


M. Keep, D.W. Haig / Tectonophysics 483 (2010) 93111

101

Fig. 7. Stratigraphic thickness versus age plots for key synorogenic sections. The Viqueque type section graph is derived from planktonic foraminiferal species data presented by Haig
and McCartain (2007); the Sabau, Mina III and Mina 1 section graphs are derived from planktonic foraminiferal and calcareous nannofossil datum levels identied by De Smet et al.
(1990). Ages of the planktonic foraminiferal datum levels used in these plots are taken from Fig. 6. Ages of the calcareous nannofossil datum levels are from Lourens et al. (2004).

absence of a major regional unconformity has been conrmed by


the detailed sampling of sections of the Batu Putih Limestone and
overlying units in West Timor by De Smet et al. (1990) and Van Marle
(1991a) and of the type section of the Viqueque Formation in East
Timor by Haig and McCartain (2007). Stratigraphic thickness versus
age plots for the sections reported by these authors are presented
in Fig. 7. These plots show a very condensed initial phase of
sedimentation followed by a rapid increase in the depositional rate
during the 32 Ma interval. This change in sedimentation rate reects
initial pelagic deposition followed by a progressive increase in the
inux of lithogenic components (to form marls initially and then
lithogenic mudstones and sandstones including turbidites). Haig and
McCartain (2007) suggested that pelagic deposition commenced
during the N18 interval (latest Miocene according to GTS2004). De
Smet et al. (1990) interpreted a N19/20 (early Pliocene according to
GTS2004) start for pelagic sedimentation. Because of the obviously
condensed nature of the pelagite section, an age difference between
the commencement of the synorogenic pelagic deposition in East
and West Timor should be viewed with caution. More stratigraphic
sections sampled at closer intervals are required to resolve this
question. In the Viqueque type section, Haig and McCartain (2007)
noted that the oldest pelagite bed with a lithogenic sand component

(viz. rare mica grains) lies near the base of zone N20 (about 4.5 Ma;
late Early Pliocene according to GTS2004) and was followed by an
increase in marl in the succession. The oldest sand bed (interpreted as
a turbidite ow) lies near the base of zone N21 (probably just younger
than 3.1 Ma; Fig. 7). This bed, interpreted to have been deposited
under middle bathyal conditions contained coastal foraminifera transported downslope. This suggested to Haig and McCartain (2007) that
islands had emerged in the Timor region by about 3.1 Ma (following
GTS2004) following the commencement of uplift at about 4.5 Ma. The
stratigraphic columns presented by De Smet et al. (1990) are not
detailed enough to determine the rst appearance of a lithogenic
component in the pelagites in the West Timor sections, but sand beds
are rst represented also within zone N21 (perhaps slightly later than
in East Timor, although this requires further study; Fig. 7).
The basis for the age determination of the Miomaffu Tuff of
the Noiltoko region has never been published (see Audley-Charles',
1986a, attribution of the age to Carter et al., 1976, where a tuff-lutite
is listed on their Table 1). Villeneuve et al. (2005, p. 305) noted that
Rosidi et al. (1979) had dated this unit as Early Miocene, although
their gures (Villeneuve et al., 2005, Figs. 8 and 9) show it as ranging
through much of the Miocene above Early Miocene oolitic limestone
of the Cablac Formation. The Bahaman facies that characterizes

Author's personal copy


102

M. Keep, D.W. Haig / Tectonophysics 483 (2010) 93111

Fig. 8. a. Aerial view of the Mt. Perdido and Mt. Laritame massifs north of the village of Ossu (see Fig. 2 for location). Red arrows point to the locations of abrupt, linear features visible
on the aerial photographs, which correspond to steep to vertical cliffs of Triassic limestone in the eld. The location of the quarry shown in Fig. 7 is outlined by a white box, as is the
location of folds in the Lari Guti Fm, which are shown in b. The line of section of e is also shown. b. Gentle folds in the Lari Guti Fm. Beds dip gently to the north and south. c. Regional
SWNE view across the Mt. Perdido and Mt. Laritame massifs, showing the Lari Guti Fm between them. d. Cross-section of the fault bounding the Lari Guti Fm to the west, indicated
by the outline box on a. Drag folds in the Lari Guti Fm terminate abruptly at the NS fault, which also appears as a lineament on aerial photos. e. View looking SW towards the Mt.
Perdido massif, with locations of probable faults creating angular topography shown by black arrows. f. Line of section across the Mt. Laritame massif, showing the striking angularity
of the range, with the locations of probable faults indicated by the black arrows.

Author's personal copy


M. Keep, D.W. Haig / Tectonophysics 483 (2010) 93111

103

Fig. 9. a and b. Fault gouge and late faults exposed in the quarry near the Mt. Perdido Massif (see Fig. 6 for location of the quarry, with and without fault annotation. See text for
discussion. The black outline box shows the location of the photograph in c. c. Close-up of minor faults in the gouge. d. Example of lineations within the quarry. See text for discussion.

the Cablac Formation has been shown to be Late Triassic or Early


Jurassic rather than Early Miocene (Haig et al., 2007), so the Noiltoko
succession requires stratigraphic revision. Consideration of the
Miomaffu Tuff in any tectonostratigraphic study should be deferred
until a detailed published analysis is made of it.
Therefore, the basis for the chronology applied by Audley-Charles
(2004), specically the absence of the N20 zone, which he took to
represent a regional unconformity and initiation of deformation in
Timor, is not viable with our interpreted chronology. There is no
published evidence for the major unconformity (including zone N20;
3.52 Ma according to the time scale he used) that suggested to him
the timing of collision. The major discordance in Timor Neogene
stratigraphy lies between the deformed indurated limestones of the
earliest Late Miocene and the friable chalk and marl of zone N18
(latest Miocene to Early Pliocene, Fig. 6).
In addition to the telescoping of Australian margin sediments, two
important processes took place during the 9.8 Ma to 5.5 Ma interval: (1)
the deformation and cementation of the Australian-Margin Megasequence pelagites; and (2) the remobilisation of Triassic to Early Jurassic
(Gondwana Megasequence) mudstones to form the Bobonaro Scaly
Clay (Audley-Charles, 1968), also referred to as the Bobonaro Melange
(Harris et al., 1998) and Synorogenic Melange (Haig et al., 2007) (Fig. 3).
Deep-sea carbonate pelagites, such as those of the AustralianMargin Megasequence on present-day Exmouth Plateau (Fig. 5b; Haq
et al., 1990) are friable chalks, marl and ooze. Similar soft sediment
would have formed the Cretaceous to early Late Miocene pelagites
of the Australian-Margin Megasequence in Timor (Haig and McCartain, 2007). Early in collision the pelagites were cemented and the
indurated limestone then cut by stylolites and calcite veins (Haig and

McCartain, 2007). Unusual faunal mixing (Haig and McCartain, 2007)


and chaotic bedding (Audley-Charles and Carter, 1972; Barber et al.,
1977), especially in the Baulaca and Cablac regions in East Timor (Haig
and McCartain, 2007, Fig. 4), are consistent with mobilization of
uids within a soft-sediment pile, and the injection of overpressured
sediment into younger horizons (Haig and McCartain, 2007).
Clasts of Cretaceous to Middle Miocene indurated pelagite with
stylolites occur in the Bobonaro Scaly Clay (= Synorogenic Melange,
Fig. 3). The basal unit of the Viqueque Formation (zone N18) lies
unconformably on the Synorogenic Melange in the Viqueque region
of East Timor (Haig and McCartain, 2007), and the lower Batu Putih
unit (belonging to zones N18N19) also unconformably overlies the
melange in West Timor. The formation of the Synorogenic Melange
therefore dates from the Late Miocene, during the 9.8 Ma5.5 Ma
interval.
Audley-Charles (1968) originally interpreted what he called the
Bobonaro Scaly Clay (Fig. 3) as olistostromal, although a melange origin
(Harris et al., 1998), especially a diapiric melange (Barber et al., 1986), is
more consistent with eld evidence in East Timor and present-day mud
volcanoes in West Timor (Barber et al., 1986; Harris et al., 1998). The
diapirism was likely the result of uid-saturated muds remobilised due
to the shortening and loading caused by collision, probably by crustal
loading due to the emplacement of the thrust nappes at this time. The
period of mud diapirism, within the 9.8 Ma to 5.5 Ma interval, therefore
implies that crustal thickening via nappe emplacement was active for
at least part of this time.
To date we have no exact timing for the emplacement of individual
thrust slices from either the Australian-derived or Banda terranes in
East Timor. However, evidence discussed in the next section suggests

Author's personal copy


104

M. Keep, D.W. Haig / Tectonophysics 483 (2010) 93111

that the main period of thrusting was during the Late Miocene.
An example of the structural complexity and deformation-timing
difculties with which we are faced, is represented in the Cablac
Mountain region of central Timor (Fig. 2), a highly deformed thrust
stack of the units belonging to the Gondwana and Australian-Margin
megasequences. This thrust stack occurs in structural contact above
the metamorphic Lolotoi Complex (Fig. 2), a thrust slice of the Banda
Terrane (Keep et al., 2009). This relationship was taken by AudleyCharles (1968) and many later authors to represent the Early Miocene
Cablac limestone (now known to be of Late Triassic to Early Jurassic
age; Haig et al., 2007) lying unconformably above the Lolotoi Complex
(now regarded as post-Jurassic in age by Harris, 2006 and Standley
and Harris, 2009). Keep et al. (2009) demonstrated that a shear zone
between the Cablac Mountain thrust stack and the Lolotoi Complex
includes sheared slivers of the upper part of the Lolotoi nappe sheet
and the lower part of the thrust stack, strongly suggesting that the
two were structurally juxtaposed. The age of the juxtaposition has
not yet been determined. Furthermore, thrust slices of Banda Terrane
material, including the Lolotoi Complex and associated younger sediments, may well be overturned in places (Keep et al., 2009). Again in
the Cablac Mountain region (Fig. 2), the Lolotoi Complex metamorphic units at the base of Cablac thrust stack occur structurally above
and in probable structural contact with relatively undeformed and
unmetamorphosed Eocene sedimentary units (e.g. zone E11, Middle
Eocene, mudstone at 9.0071S, 125.6328E), also of the Banda Terrane
(Haig and McCartain, 2007; Keep et al., 2009). Given that both
lithologies derive from the Banda Terrane, their eld relationships
indicate the possibility that the Banda Terrane nappes may be overturned in places. Although overturned stratigraphy is common in
thrust nappes, many correlations from West Timor into East Timor
assume that the stratigraphy is continuous, not overturned (e.g.
Harris, 2006; Standley and Harris, 2009). If the Banda Terrane nappe
in the Cablac region is overturned (Keep et al., 2009), the relationship
with the overlying thrust stack of Gondwana material is still not clear.
The overturning may have occurred prior to the emplacement of the
Gondwana material structurally above.
5.3. Tectonic quiet zone
From 5.5 to 4.5 Ma carbonate pelagites of the lower to middle
bathyal zone were deposited over much of Timor (Haig and McCartain,
2007). These rocks are friable chalk and marl and lie unconformably
above the Synorogenic (Bobonaro) Melange, as the Batu Putih Member
of the Viqueque Formation (Synorogenic Megasequence) (Haig and
McCartain, 2007) (Fig. 2). The extent of these undeformed pelagites
(Audley-Charles 1968, Kenyon, 1974, Haig and McCartain, 2007) implies that the tectonic conditions that existed during their deposition
extended across much of Timor, having a regional and not local cause.
If, as the deposition of the pelagic sediments implies, we have
a period of about 1 Ma of undisturbed lower to middle bathyal
conditions (probably at water depths of 1500 to 2500 m), then this
represents a critical stage in the development of the young orogen, as
it is evidence of a tectonic hiatus in the middle of ongoing collision.
Although deposition of sediments within deforming orogenic piles is
known (e.g. syn-tectonic basins in an accretionary prism, e.g. Twiss
and Moores, 2007), these sediments are usually rapidly deformed
during continuing orogenesis.
The deformed Siu Formation in the Kolbano area of West Timor,
discussed above, has previously been interpreted as including Early
Pliocene beds coeval with the basal part of the undeformed Synorogenic
Megasequence, implying at least the basal part of the Synorogenc
Megasequence is deformed in this region. Sawyer et al. (1993, p. 543)
suggested that the gently deformed Viqueque Formation may have
acted as a passive roof thrust over the Kolbano Sequence (in which
they included the deformed Siu Formation, then referred to the Ofu
Formation). As pointed out above, no evidence for deposits belonging to

Late Miocene zones N16 and N17 has been published for the Siu
Formation, and an unconformity, similar to that found in the East Timor,
incorporating this interval might be present. The latest Miocene to Early
Pliocene pelagites of the Siu Formation may represent the product of
local deformation in the Kolbano region, or the age may have been
misidentied from difcult thin section analysis. From a regional
perspective, if problems with the Siu Formation and the Miomaffu Tuff
(also discussed above) are left aside until further study, Pliocene units in
Timor are relatively undeformed, and only pre-Late Miocene units are
present in the thrust stacks. This implies that the main period of
thrusting was during the Late Miocene.
We propose that the quiet interval (5.54.5 Ma) may well represent
the time when subduction became locked by the great thickness of
continentplateau crust entering the subduction system. AudleyCharles (2004) placed this interval at 3.52 Ma immediately after his
assumed N20 unconformity. It is signicant that very little contemporaneous volcanic material is present in the Pliocene sediments of Timor.
Kenyon (1974) identied only rare thin tuffaceous beds near the base
(zone N18N19) of the Batu Putih Limestone in West Timor.
Haig and McCartain (2007) pointed out that over much of Timor
during this quiet interval water depths were probably in the range of
1500 m to 2500 m, similar to water depths interpreted for the precollisional continental plateau. Crustal loading by the thrust pile that
developed during the Late Miocene may have kept water depths over
much of the Timor region within this lowermiddle bathyal zone for
1 Ma after the main period of thrusting and locking of the subduction
system. The continuation of volcanic activity on the Inner Banda
Arc adjacent Timor until about 3 Ma (Honthaas et al., 1998) may
represent a lag effect.
5.4. Emergence and late extension
We suggest that the emergence and modern topography of Timor
was associated with isostatic rebound of the thrust pile that was
negatively buoyant during the Early Pliocene due to a combination of
crustal loading and the detaching subducting slab during the Late
Miocene (e.g. Kaneko et al., 2007). Van Marle (1991a), De Smet et al.
(1990) and Haig and McCartain (2007) showed that within the
Synorogenic Megasequence on the southern side of Timor there was
transition in the depositional environments from the lower part of the
middle bathyal zone to the upper part of the middle bathyal zone
during the Middle to Late Pliocene, and outlined evidence for the
uplift and emergence of the northern part of Timor during this time.
As indicated above, the lowest occurrence of rare mica of sand size in
lowermiddle bathyal marls deposited at about 4.5 Ma is the rst
evidence of erosion of pre-Neogene units incorporated in the Timor
Orogen that we have found in the Synorogenic Megasequence (Haig and
McCartain, 2007). Haig and McCartain (2007) showed that the oldest
evidence for the emergence of an island in the East Timor region is
recorded in turbidite beds, deposited in the middle bathyal zone, that
occur low in zone N21 (probably close to 3.1 Ma). These beds include
contemporaneous biogenic material derived from the coastal zone as
well as an abundance of phyllite clasts. Similar turbidites belonging to
zone N21 also rst appear at this level in the Pliocene succession of West
Timor (De Smet et al., 1990; van Marle, 1991). These turbidites mark the
onset of a signicant change in the sedimentation pattern, indicative of
tectonic uplift. Higher turbidites contain clasts representing a diverse
range of rock units derived from the Gondwana and Australian-Margin
megasequences as well as the Banda Terrane.
The total surface uplift in Timor from an Early Pliocene seaoor at
about 2000 m water depth to present-day mountains approaching
3 km above sea level (which are erosional remnants) exceeds 5 km.
Audley-Charles (2004) noted the very steep slope on the northern side
of Timor facing the 3 km-deep Wetar Strait (Fig. 2), and suggested that
this was a major fault, that he named the Wetar Suture, attributed to
the Pliocene. According to Audley-Charles (2004), the Wetar Suture

Author's personal copy


M. Keep, D.W. Haig / Tectonophysics 483 (2010) 93111

separates Australian continental cover rocks from subsided Banda


forearc (forming Wetar Strait), and this was evidence that the
Australian-derived units had overridden the Banda Trench. Grady
(1975) and Grady and Berry (1977) suggested that high-angled faults
played a signicant role in Timor uplift, and Audley-Charles (2004)
also noted the importance of regional uplift, although his timing of this
cannot be supported (see above), and suggested a possible isostatic
response to the locking of subduction. Mobilization of the Synorogenic
Melange continued during the uplift phase and is evidenced today
by mud volcanoes in West Timor (Barber et al., 1986).
An example of a late-stage high-angle fault that currently forms a
striking topographic feature in East Timor is the northern slope of
Cablac Mountain (Haig et al., 2008; Keep et al., 2009) where an
extensive crush breccia has developed. In the Ossu region (Fig. 2) of
East Timor, we observed, at the largest scale, late-stage faults that are
dominant linear features, visible from aerial and satellite images,
which dissect prominent limestone massifs and coincide with steep
limestone cliffs (Fig. 8). This can clearly be seen near the village of
Ossu, north of Viqueque (Figs. 2, 8), where prominent linear features
truncate the Mt. Perdido and Mt. Laritame Triassic/Early Jurassic
limestone massifs (Fig. 8). These lineaments on aerial photos
correspond to steep (N70) to vertical cliffs of limestone, which dip
between 75 and 90, and which dominate the highest topography of
the area. The angular nature of the highest peaks is also suggestive of
structural control (Fig. 8e). At a smaller scale, a quarry on the road
north of Ossu village (at 8.71173S, 126.364E), along strike from
linear features seen on aerial photos, exposes a fault zone of highlysheared and strongly-lineated Late Triassic or Early Jurassic limestones of the Gondwana Megasequence (Fig. 9). This complex array
of faults indicates signicant offset, and the thickness of the gouge
(over 30 m of structural thickness is exposed in the quarry) attests
to a prolonged period of brittle shear.
Further examples of young deformation occur within Pliocene
Pleistocene units of the Synorogenic Megasequence between Ossu and
Viqueque (Fig. 2), close to the fault quarry (Fig. 8). Gentle folding in
rocks of the latest Pliocene or earliest Pleistocene Lari Guti Formation
(at 8.70495S, 126.36222E) displays dips of 10 north and south about
a gentle EW striking fold axis (Fig. 6b). This gentle dip, in very young
units in close proximity to the complex array of faults exposed in the
quarry (Fig. 9), is suggestive of block rotations during faulting. Further south at the Cuha River in Viqueque (Fig. 10), late listric normal
faults cut the Pliocene Viqueque Formation (Fig. 10) and dip gently
to the northeast (151/21NE). Rotation on these listric faults is
consistent with the gentle (25) south to southwesterly dip of the
entire Viqueque Formation (065/23SE) at this location.

105

Evidence for late and probable ongoing extensional/transtensional


deformation is compelling, and additional eldwork will certainly
document more examples at various scales across the island. We suggest
this late extension may play an important part in the exhumation of the
thrust nappes, and may account for much of the present-day topography
of East Timor.
6. Discussion
We propose that the orogenic cycle as recorded in East Timor can be
divided into three distinct phases, as detailed above. The timing of our
proposed phases is based on detailed biostratigraphic determinations
across East Timor, but this must be reconciled with other known ages
for tectonic events in the region. The timing of collision in particular
has implications for tectonic models of the Banda Arc in general.
We suggest that initial collision occurred at sometime during the
Late Miocene between 9.8 Ma and 5.5 Ma, which is signicantly older
than some other estimates for collision (e.g. 3 to 4 Ma, Hall, 2002;
Audley-Charles, 2004; Roosmawati and Harris, 2009), but is consistent with other evidence from both Timor and the adjacent Australian
North West Shelf (e.g. around 8 Ma, Reed, 1985; McCaffrey et al.,
1985; Berry and McDougall, 1986; Richardson and Blundell, 1996;
Charlton, 2000; Rutherford et al., 2001; Keep et al., 2002; Keep et al.,
2003; Haig and McCartain, 2007). One of the main arguments for a
collision at 3 to 4 Ma, is that the youngest volcanism recorded in this
part of the Banda Arc is only 11 Ma (e.g. Hartono, 1990; Hall, 2002),
and continues until 3 Ma in places (Honthaas et al., 1998). It therefore
seems difcult to have a collision so soon after the initiation of the arc.
Also, Hall (2002; Fig. 23) shows that at our proposed time of collision
post 9.8 Ma, that the subducting part of the arc was approximately
600 km to the north, a distance underlain by oceanic lithosphere,
implying that collision could not have occurred as continental crust
was still a signicant distance from the arc at that time.
However, there are a number of lines of evidence that point to
an earlier age of collision in this region. Firstly, on Timor, the Aileu
Metamorphic Complex (part of the Gondwana Megasquence, Figs. 2
and 3) records 40Ar/39Ar and K/Ar ages for hornblendes of 8 Ma (Berry
and Grady, 1981). These ages indicate that retrograde metamorphism
occurred along this part of the Australian margin at this time, presumably in response to a tectonic stimulus, implying that the prograde
phase (collision?) occurred sometime prior to 8 Ma. That stimulus may
be the same one responsible for deformation across the southerlyadjacent Timor Sea and Browse Basin (Fig. 2) at approximately the same
time. Seismic structural analyses on both regional 2D and local 3D
seismic data across the Timor Sea and Browse Basin indicate a period of

Fig. 10. a. Aerial view of the Cuha River in the town of Viqueque. The contact between the Viqueque Fm and the Bobonaro Melange occurs in the river, and is indicated by a red dashed
line. The location of the listric faults is indicated by a red dot, in a small tributary to the Cuha River close to the conact. b. NE-dipping listric faults in the Viqueque Fm. Faults strike to
the SE and dip moderately to gently to the NE.

Author's personal copy


106

M. Keep, D.W. Haig / Tectonophysics 483 (2010) 93111

regional tectonic activity at 8 Ma (Baillie and Jacobson, 1995; Keep et al.,


2002; Harroweld et al., 2003; Harroweld and Keep, 2005). This
requires at least a localised tectonic event at around 8 Ma in the area
immediately adjacent to present-day Timor.
The lateral extent of the proposed circa 8 Ma deformation may also
be recorded in Sumba to the west (Fig. 2). The issue of the origin of
Sumba and how it came to be in its current position has been the
subject of some discussion (e.g. Audley-Charles, 1985; van der Werff
et al., 1994; Wensink, 1994, 1997; van der Werff, 1995; Wensink and
van Bergen, 1995; Fortuin et al., 1994, 1997; Soeria-Atmadja et al.,
1998; Abdullah et al., 2000; Rutherford et al., 2001; Lytwyn et al.,
2001). However, there seems to be general agreement that uplift of
Sumba started during the Late Miocene, from approximately 7 Ma
(van der Werff et al., 1994; Fortuin et al., 1994, 1997; Rutherford et al.,
2001), although its exact location at this time is disputed (e.g.
compare Rutherford et al., 2001 with Hall, 2002). The coincidence of
the start of uplift of Sumba with the timing of deformation on the
North West Shelf led Keep et al. (2003), using the location of Sumba as
proposed by Rutherford et al. (2001) to propose that the uplift of
Sumba was the result of collision with an outlying part of Australian
continental crust, similar to the present-day geometry of the Exmouth
Plateau (Keep et al., 2003, Fig. 10), a view more recently echoed by
Roosmawati and Harris (2009, Fig. 12) who show a remarkably
similar geometry of Australian continental crust underlying Sumba.
However, the idea of Australian continental crust underlying Sumba
is not new. Chamalaun (1982b) on the basis of Bouguer gravity,
suggested that Sumba was underlain by continental crust at least
24 km thick, a fact used by Hartono (1990) in his illustration of the
tectonic development of the Banda Sea area, and illustrated by van der
Werff (1995; Fig. 10), who clearly shows continental crust beneath
Sumba. As Sumba is thought to be a fragment of arc (e.g. van der Werff
et al., 1994; Fortuin et al., 1994, 1997; Rutherford et al., 2001), the
continental crust presumably results from underplating from collision. Furthermore, Chamalaun et al. (1982a) indicated that the south
part of the Savu Sea was also underlain by continental crust, possibly
old, thinned passive margin material. Interpreting acoustic basement
structures between Sumba and Timor, van der Werff et al. (1994) and
van der Werff (1995) proposed structural continuity of both Sumba
and Timor since the late Miocene, as did Karig et al. (1987), citing in
particular the continuity of a series of sub-sea ridges and back-thrusts
local to the Sumba area (see van der Werff, 1995: Fig. 1), an interpretation consistent with the Chamalaun et al. (1982a) determination of
continental crust in the region. In cartoons depicting the tectonic
evolution of Sumba, Abdullah et al. (2000; Fig. 7d) also seemed to imply
the encroachment of Australian continental crust immediately adjacent
to Sumba.
There therefore seems to be some body of evidence for collision
occurring earlier than 34 Ma in the Timor region, possibly affecting
Sumba as well as Timor. For collision to have occurred earlier, either
some segment of the arc was further to the south than shown in
current reconstructions (Hall, 2002), or continental crust extended
further to the north, or both.
If the irregular nature of the colliding margin is considered (Fig. 4),
it is possible that continental crust extended from 500 to 900 km
further out towards the subduction zone than currently accepted
(Hall, 2002), as a continental plateau (the Timor Plateau of Haig and
McCartain, 2007). The entire distance to the arc could have been
occupied by continental rather than oceanic lithosphere (e.g.
McCaffrey et al., 1985; also compare Fig 4 to Hall, 2002; Fig. 23),
and that collision could therefore have occurred prior to 5 Ma. Various
other lines of evidence also suggest some segment of the arc in the
Timor/Sumba region may have been further south at around 10 to
8 Ma. Van der Werff et al. (1994) suggested that outcrops of Early
Miocene volcanic rocks on West Sumba, Miocene volcaniclastic rocks
on East Sumba, and evidence that the Savu Basin lled from the south
indicate that a proto-arc must have existed south of the Sumba

Ridge in the early, and possibly Mid- to Late Miocene. A southerly


source for volcaniclastic turbidites on Sumba, rather than the volcanic
arc to the north, was also the surprising conclusion for Fortuin et al.
(1997). Van der Werff et al. (1994) and Wensink and van Bergen
(1995) therefore proposed that a small arc may have been active
south of Sumba from the Early Miocene until the Late Miocene, and
that it is probably now buried beneath the encroaching deformation
front. These authors further suggested that such a change in the
location of the volcanic front from that currently predicted (e.g. Hall,
2002), may be related to changes in subduction geometry as a result of
the shape of the Australian rifted margin. Wensink and van Bergen
(1995) also propose that arrival of buoyant crustal materials at
the southerly proto arc would have caused subduction dip to atten,
as observed in the region by McCaffrey (1989), causing the arctrench
gap to widen, thus shifting the arc segment north, closer to its predicted position. It therefore seems plausible that an earlier collision
along one section of the arc only, could have been possible.
The main argument against an earlier age of collision is the
continuation of volcanism until around 3 Ma in the Wetar region
(Honthaas et al., 1998), leaving an apparent gap of at least 6 Ma from
our proposed onset of collision until cessation of volcanism. However,
our proposed model for collision is episodic, with nal locking of the
subduction zone occurring only in the 5.5 Ma to 4.5 Ma interval.
Assuming some degree of slab melting would be possible until the
subduction zone is fully jammed, the discrepancy in timing decreases
to only 1.5 Ma, which may simply represent a lag effect. Volcanism at
Wetar is also episodic in nature; Honthaas et al. (1998) recorded three
periods of volcanic activity around Wetar, with ages of 7.76.5 Ma,
5 Ma and 43 Ma. Samples from the Wetar segment showed larger
isotopic variations than any of the other areas sampled by Honthaas
et al. (1998), and indicate a strong involvement of Australian continental crust (Honthaas et al., 1998). However, assuming that the
Australian margin was located considerably further to the south than
we are currently proposing, Honthaas et al. (1998) ruled out the
possibility that the isotopic signatures they encountered near Wetar
represented the subduction of the continental margin or proximal
sediments. However, it appears to be the case that tectonic models
proposed for the collision along this segment of the Banda Arc (for
example Hall, 2002; Audley-Charles, 2004; Roosmawati and Harris,
2009) have assumed a uniform down-going slab, without any slab
rupture. Vertical or horizontal slab rupture may occur commonly
during subduction (Nolet, 2009), and has been clearly documented
at other subduction margins (Obayashi et al., 2009). A vertical or
horizontal tear in the down-going slab would help to explain both
magmatic and vertical motions at the surface (Nolet, 2009), including,
for Timor, the late-stage volcanism recorded at Wetar and potentially
even part of the uplift observed in Timor (Kaneko et al., 2007). Again
the idea of slab rupture in the area is not new. McCaffrey et al. (1985)
proposed both rupture and local tearing beneath Timor to explain the
distribution of microseismicity data. They proposed that the slab was
detaching along a steep fault parallel to the northwest coast of Timor,
and further proposed a tear decoupling seismic zones across Timor
Island (McCaffrey et al., 1985). Such rupture of the down-going slab
could well have lead to ongoing, episodic volcanism around Wetar,
allowing an inux of hot material from beneath the rupturing slab.
The coincidence of timing between magmatic events on Wetar (7.7
6.5 Ma, 5 Ma, 43 Ma) and the tectonic intervals we propose on Timor
(post 9.8 Ma collision, 9.8 to 5.5 Ma shortening, 5.5 to 4.5 Ma quiet
zone) may be indicative of a causal relationship or of discrete episodes
of slab rupture.
A possible earlier age for collision in the Timor/Sumba region, at
around 8 Ma rather than 3 Ma, need not have major implications for
regional tectonic models, such as those of Hall (2002) and in fact
requires only small changes that have limited effects. We suggest that
this earlier phase of collision of an outlying Timor Plateau would
have affected approximately 1000 km of the margin only, equivalent

Author's personal copy


M. Keep, D.W. Haig / Tectonophysics 483 (2010) 93111

to the proposed width of the plateau, modelled on the present-day


Exmouth Plateau (slightly wider if an equivalent Wallaby Plateau
geometry is included, Fig. 11). We propose that the segment including
the protruding plateau arrived at the subduction zone some time after
9.8 Ma, but well before 5.5 Ma (Fig. 11a). Fig. 11 also shows the
notional geometry of the Wallaby Plateau, which in the model would
have collided slightly earlier than the main plateau. Segments to the
east of the proposed Timor Plateau would have been embayments
(Fig. 11a), which would not yet have collided. The period from 9.8 to
5.5 Ma involved shortening and deformation of the collided plateau,
involving the shaded areas of Fig. 11b. According to Keep et al. (2003)
parts of the Australian continental crust started to underplate Sumba
around this time, coincident with the onset of uplift of Sumba Island,
which may have been caused by a Wallaby Plateau-type outlier
colliding with Sumba at this time (Fig. 11b). By 5.5 Ma the plateau had

107

completely jammed subduction, resulting in a temporary cessation of


shortening and the development of widespread pelagic conditions
across Timor (Fig. 11c). We propose a period of tectonic quiescence
after the initiation of collision, during which lower to middle bathyal
carbonates were deposited over much of Timor. This would be a
critical stage in the developing orogeny, which would not usually be
preserved in older orogens. We propose that this quiet interval (5.5
4.5 Ma) represents the locking of subduction as a result of continental
crust entering the subduction system, allowing deep-water (1500 m
to 2500 m) to prevail over much of Timor at this time. Uplift of
Sumba was well underway by this time. The period from 4.5 Ma to
present was dominated by the rapid rise of Timor as a result of slab
detachment (e.g. Kaneko et al., 2007). The segmentation of the margin
means that some areas of embayments, such as that to the immediate
east of Timor Plateau, would only have collided post 4.5 Ma, with

Fig. 11. Schematic tectonic model for the stages of collision in the Timor region. The solid black line shows the proposed geometry of the pre-collisional Timor Plateau, as shown in
Fig. 4a. The dashed line shows the additional complications caused by a pre-collisional geometry more similar to that shown in Fig. 4b, including the Wallaby Plateau. a. Initial
collision occurred immediately post 9.8 Ma, when a salient of the Australian margin, the Timor Plateau, rst collided with the arc, along a limited segment of the margin. Adjacent
embayments would not collide until post 4.5 Ma. It is possible that some section of the segmented margin may have collided rst (dashed line) to the west. b. Between 9.8 Ma and
5.5 Ma the collided segment(s) of the Australian margin shortened (areas shown as shaded). This shortening was accompanied by deformation and cementation of the Australian
Margin Megasequence, and the remobilization of Gondwana Megasequence muds. The uplift of Sumba commenced from approximately 7 Ma (see text for discussion), coincident
with the shortening of a Wallaby Plateaux-type geometry at this time. c. From 5.5 Ma to 4.5 Ma pelagic conditions prevailed across Timor as a result of the Timor Plateau jamming the
suduction zone at this time (see shaded areas for approximate areas that would have been shortened). d. From 4.5 Ma onwards, the evolution of Timor was dominated by uplift from
detachment of the downgoing slab.

Author's personal copy


108

M. Keep, D.W. Haig / Tectonophysics 483 (2010) 93111

corresponding development of sea oor spreading lineations in the


eastern part of the Southern Banda Sea (Hartono, 1990; Hinschberger
et al., 2005).
Acknowledgements
We acknowledge the University of Western Australia for continuing support for our studies in East Timor. It is with great appreciation
that we thank ENI Pty Ltd for recognizing the importance of our
studies and recently making funding available to us. During eld work
in 2007 and 2008 we beneted greatly from stimulating discussions
of Eujay McCartain and Ken McCormack from the University of
Western Australia, Mark Quigley from the Canterbury University,
and Patricia Rich from Monash University. We thank Alfredo Pires,
Secretary of State for Natural Resources of Timor Leste, Francisco da
Costa Monteiro and the staff of the Secretariate for Energy and Natural
Resources (SERN), and Gualdino da Silva and his staff at the Autoridade Nacional de Petroleo in Dili for their continuing support. We
thank Robert Hall, Nadine McQuarrie, Mark Quigley and Wouter
Schellart for reviews, comments and observations that helped us to
improve the manuscript.
References
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 magmatic rocks. Journal of Asian Earth Sciences 18, 533546.
Audley-Charles, M.G., 1968. The geology of Portuguese Timor. Geological Society of
London, Memoir, vol. 4.
Audley-Charles, M.G., 1981. Geometrical problems and implications of large scale
overthrusting in the Banda ArcAustralian margin collision zone. In: McClay, K.R.
(Ed.), Thrust and Nappe Tectonics: Geological Society of London Special Publication,
vol. 9, pp. 407416.
Audley-Charles, M.G., 1983. Reconstruction of eastern Gondwanaland. Nature 306, 4850.
Audley-Charles, M.G., 1985. The Sumba enigma: is Sumba a diapiric fore-arc nappe in
process of formation? Tectonophysics 119, 435449.
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, 161175.
Audley-Charles, M.G., 1986b. TimorTanimbar Trough: the foreland basin of the
evolving Banda orogen. Special Publication of the Association of Sedimentologists,
vol. 8, pp. 91102.
Audley-Charles, M.G., 1988. Evolution of the southern margin of Tethys (North Australian
region) from Early Permian to Late Cretaceous. Geological Society Special Publications,
vol. 37, pp. 79100.
Audley-Charles, M.G., 2004. Ocean trench blocked and obliterated by Banda forearc
collision with Australian proximal continental slope. Tectonophysics 389, 6579.
Audley-Charles, M.G., Carter, D.J., 1972. Paleogeographical signicance of some aspects
of Palaeogene and Early Neogene stratigraphy and tectonics of the Timor Sea
region. Palaeogeography, Palaeoclimatology, Palaeoecology 11, 247264.
Audley-Charles, M.G., Harris, R.A., 1990. Allochthonous terranes of the southwest
Pacic and Indonesia. Philosophical Transactions of the Royal Society London A331,
571587.
Audley-Charles, M.G., Carter, D.G., Barber, A.J., 1974. Stratigraphic basis for tectonic
interpretations of the Outer Banda Arc, Eastern Indonesia. Proceedings Indonesian
Petroleum Association, Third Annual Convention, June 1974, pp. 2544.
Audley-Charles, M.G., Carter, D.J., Barber, A.J., Norvick, M.S., Tjokrosapoetro, S., 1979.
Reinterpretation of the geology of Seram: implications for the Banda Arcs and
northern Australia. Journal of the Geological Society of London 136, 547568.
Audley-Charles, M.G., Ballantyne, P.D., Hall, R., 1988. MesozoicCenozoic rift-drift
sequence of Asian fragments from Gondwanaland. Tectonophysics 155, 317330.
Bachri, S., Situmorang, R.L., 1994. Geological Map of the Dili Quadrangle 24062407, Scale
1:250,000. Geological Research and Development Centre, Bandung, Indonesia.
Baillie, P.W., Jacobson, E., 1995. Structural evolution of the Carnarvon Terrace, Western
Australia. APEA Journal 35, 321332.
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,
119128.
Barber, A.J., Audley-Charles, M.G., Carter, D.J., 1977. Thrust tectonics in Timor. Journal of
the Geological Society of Australia 24, 5162.
Barber, A.J., Tjokrosapoetro, S., Charlton, T.R., 1986. Mud volcanoes, shale diapirs, wrench
faults, and melanges in accretionary complexes, eastern Indonesia. American
Association of Petroleum Geologists Bulletin 70, 17291741.
Barkham, S.T., 1993. The structure and stratigraphy of the Permo-Triassic carbonate
formations of West Timor, Indonesia. University of London, unpublished PhD thesis,
397 pp.
Berry, R.F., 1981. Petrology of the Hili Manu lherzolite, East Timor. Journal of the
Geological Society of Australia 28, 453469.

Berry, R.F., Grady, A.E., 1981. Deformation and metamorphism of the Aileu Formation,
north coast, East Timor and its tectonic signicance. Journal of Structural Geology 3,
143167.
Berry, R.F., Jenner, G.A., 1982. Basalt geochemistry as a test of the tectonic models of
Timor. Journal of the Geological Society of London 139, 593604.
Berry, R.F., McDougall, I., 1986. Interpretation of 40Ar/39Ar and K/Ar dating evidence
from the Aileu Formation, East Timor, Indonesia. Chemical Geology 59, 4358.
Beyrich, E., 1865. ber eine Kohlenkalk-Fauna von Timor. Abhandlungen der
Koniglichen Akademie der Wissenschaften zu Berlin 1864, 5998.
Bird, P.R., 1987. The geology of the Permo-Trias of Kekneno, West Timor. University of
London, unpublished PhD thesis, 265 pp.
Bird, P.R., Cook, S.E., 1991. Permo-Triassic successions of the Kekneno area, West Timor:
implications for palaeogeography and basin evolution. Journal of Southeast Asian
Earth Sciences 6, 359371.
Blow, W.H., 1969. Late middle Eocene to Recent planktonic foraminiferal biostratigraphy. In: Brill Leiden, E.J. (Ed.), International Conference on Planktonic Microfossils 1st Geneva 1967 Proceedings, vol. 1, pp. 199421.
Bowin, C., Purdy, G.M., Johnston, C., Shor, G., Lawver, L., Hartono, H.M.S., Jezek, P., 1980.
Arccontinent collision in the Banda Sea region. American Association of Petroleum
Geologists Bulletin 64, 868915.
Brnnimann, P., Resig, J., 1971. A Neogene globigerinacean biochronologic time-scale
of the south-western Pacic. Initial Reports of the Deep Sea Drilling Project, vol. 7,
pp. 12351469.
Brouwer, H.A., 1942. Summary of the geological results of the expedition. In: Brouwer,
H.A. (Ed.), Geological Expedition of the University of Amsterdam to the Lesser
Sunda Islands in the South Eastern Part of the Netherlands East Indies 1937, vol. IV.
N.V. Noord-Hollandsche Uitgevers Maatschappij, Amsterdam, pp. 345401.
Browne, M., Earle, M.M., 1983. Cordierite-bearing schists and gneisses from Timor,
eastern Indonesia: PT conditions of metamorphism and tectonic implications.
Journal of Metamorphic Geology 1, 183203.
Brunnschweiler, R.O., 1978. Notes on the geology of Eastern Timor. In: Belford, D.J.
(Ed.), The Crespin Volume: Essays in Honour of Irene Crespin. In: Scheibnerova
(Ed.), Bureau of Mineral Resources Australia Bulletin, vol. 192, pp. 918.
Carter, D.J., Audley-Charles, M.G., Barber, A.J., 1976. Stratigraphical analysis of island
arccontinental margin collision in eastern Indonesia. Journal of the Geological
Society of London 132, 179198.
Chamalaun, F.H., 1977. Paleomagnetic evidence for the relative positions of Timor and
Australia in the Permian. Earth and Planetary Science Letters 34, 107112.
Chamalaun, F.H., Grady, A.E., 1978. The tectonic development of Timor: a new model
and its implications for petroleum exploration. APEA Journal 18, 102108.
Chamalaun, F.H., Lockwood, K., White, A., 1976. The Bouguer gravity eld and crustal
structure of Eastern Timor. Tectonophysics 30, 241259.
Chamalaun, F.H., Grady, A.E., von der Borch, C.C., Hartono, H.M.S., 1982a. The signicance
of Sumba Island (Indonesia). American Association of Petroleum Geologists Bulletin
361375.
Chamalaun, F.H., Grady, A.E., von der Borsch, C.C., Hartono, H.M.S., 1982b. Banda Arc
tectonics: the signicance of the Sumba Island (Indonesia). American Association
of Petroleum Geologists, Memoir 34, 261375.
Chang, C.-P., Chang, T.-Y., Angelier, J., Kao, H., Lee, J.-C., 7Yu, S.-B., 2003. Strain and stress
eld in Taiwan oblique convergent system: constraints from GPS observations and
tectonic data. Earth and Planetary Science Letters 214, 115127.
Charlton, T.R., 1987. The tectonic evolution of the KolbanoTimor trough accretionary
complex, Timor, Indonesia. Unpublished PhD thesis, University of London, 374 pp.
Charlton, T.R., 1989. Stratigraphic correlation across an arccontinent collision zone:
Timor and the Australian Northwest Shelf. Australian Journal of Earth Sciences 36,
263274.
Charlton, T.R., 1991. Postcollisional extension in arccontinent collision zones, eastern
Indonesia. Geology 19, 2831.
Charlton, T.R., 2000. Tertiary evolution of the eastern Indonesia collision complex.
Journal of Asian Earth Sciences 18, 603631.
Charlton, T.R., 2002. The structural setting and tectonic signicance of the Lolotoi,
Laclubar and Aileu metamorphic massifs, East Timor. Journal of Asian Earth Sciences
20, 851865.
Charlton, T.R., Wall, D., 1994. New biostratigraphic results from the Kolbano area,
southern West Timor: implications for the MesozoicTertiary stratigraphy of Timor.
Journal of Southeast Asian Earth Sciences 9, 113122.
Cook, S.E., 1986. Triassic sediments from East Kekneno, West Timor. University of
London, unpublished PhD thesis, 384 pp.
Crostella, A.A., Powell, D.E., 1975. Geology and hydrocarbon prospects of the Timor area.
Proceedings of the Fourth Annual Convention. Indonesian Petroleum Association,
Jakarta. June 1975, 17 pp.
Daly, M.C., Cooper, M.A., Wilson, I., Smith, D.G., Hooper, B.G.D., 1991. Cenozoic plate
tectonics and basin evolution in Indonesia. Marine and Petroleum Geology 8, 221.
De Bruyne, D.L., 1941. Sur la composition et la gense du Bassin Central de Timor. In:
Brouwer, H.A. (Ed.), Geological Expedition of the University of Amsterdam to the
Lesser Sunda Islands in the South Eastern Part of the Netherlands East Indies 1937,
vol. 3. N.V. Noord-Hollandsche Uitgevers Maatschappij, Amsterdam. 238 pp.
De Roever, W.P., 1940. Geological investigations in the southwestern Moetis Region
(Netherlands Timor). In: Brouwer, H.A. (Ed.), Geological Expedition of the
University of Amsterdam to the Lesser Sunda Islands in the South Eastern Part
of the Netherlands East Indies 1937, vol. 2. N.V. Noord-Hollandsche Uitgevers
Maatschappij, Amsterdam. 101 pp.
De Smet, M.E.M., Fortuin, A.R., Troelstra, S.R., Van Marle, L.J., Karmini, M., Tjokrosapoetro, S., Hadiwasastra, S., 1990. Detection of collision-related vertical movements in
the Outer Banda Arc (Timor, Indonesia) using micropalaeontological data. Journal
of Southeast Asian Earth Sciences 4, 337356.

Author's personal copy


M. Keep, D.W. Haig / Tectonophysics 483 (2010) 93111
De Waard, D., 1954a. Contributions to the geology of Timor: I. Geological research in
Timor, an introduction. Madjalah Ilmu Alam Untuk Indonesia (Indonesian Journal
for Natural Science) 110, 18.
De Waard, D., 1954b. Contributions to the geology of Timor: II. The orogenic main phase
in Timor. Madjalah Ilmu Alam Untuk Indonesia (Indonesian Journal for Natural
Science) 110, 920.
De Waard, D., 1954c. Contributions to the geology of Timor: III. Structural development
of the crystalline schists in Timor. Tectonics of the Lalan Asu massif. Madjalah Ilmu
Alam Untuk Indonesia (Indonesian Journal for Natural Science) 110, 143153.
De Waard, D., 1954d. Contributions to the geology of Timor: VI. The second geological
Timor expedition, preliminary results. Madjalah Ilmu Alam Untuk Indonesia
(Indonesian Journal for Natural Science) 110, 154160.
De Waard, D., 1955a. Contributions to the geology of Timor: VII. On the tectonics of the
Ofu Series. Madjalah Ilmu Alam Untuk Indonesia (Indonesian Journal for Natural
Science) 111, 137143.
De Waard, D., 1955b. Contributions to the geology of Timor: VIII. Tectonics of the
Sonnebait overthrust unit near Nikiniki and Basleo. Madjalah Ilmu Alam Untuk
Indonesia (Indonesian Journal for Natural Science) 111, 144150.
De Waard, D., 1956. Contributions to the geology of Timor: IX. Geology of a NS section
across western Timor. Madjalah Ilmu Alam Untuk Indonesia (Indonesian Journal
for Natural Science) 112, 101113.
De Waard, D., 1957. Contributions to the geology of Timor: XII. The third Timor
geological expedition, preliminary results. Madjalah Ilmu Alam Untuk Indonesia
(Indonesian Journal for Natural Science) 113, 742.
Direen, N.G., Stagg, H.M.J., Symonds, P.A., Colwell, J.B., 2008. Architecture of volcanic
rifted margins: mew insights from the Exmouth-Gascoyne margin, Western
Australia. Australian Journal of Earth Sciences 55, 341363.
Earle, M.M., 1979. Mesozoic ophiolite and the blue amphibole on Timor and the
dispersal of eastern Gondwanaland. Nature 282, 375378.
Exon, N.F., Haq, B.U., von Rad, U., 1992. Exmouth Plateau revisited: scientic drilling and
geological framework. Proceedings of the Ocean Drilling Program, Scientic Results
122, 320.
Falloon, T.J., Berry, R.F., Robinson, P., Stolz, A.J., 2006. Whole-rock geochemistry of the
Hili Manu peridotite, East Timor: implications for the origin of Timor ophiolites.
Australian Journal of Earth Sciences 53, 637649.
Fortuin, A.R., Roep, Th.B., Sumosusastro, P.A., 1994. The Neogene sediments of east
Sumba, Indonesia products of a lost arc? Journal of Southeast Asian Earth Sciences
9, 6779.
Fortuin, A.R., Van der Werff, W., Wensink, H., 1997. Neogene basin history and
palaeomagnetism of a rifted and inverted forearc region, on- and offhsore Sumba,
Eastern Indonesia. Journal of Asian Earth Sciences 15, 6188.
Gageonnet, R., Lemoine, M., 1957a. Sur la stratigraphie de l'autochtone au Timor
Portugais. Les Comptes rendus de l'Academie des sciences 244, 21682171.
Gageonnet, R., Lemoine, M., 1957b. Composition et subdivisions du complex charrie
au Timor Portugais. Les Comptes rendus de l'Academie des sciences 244,
22462249.
Gageonnet, R., Lemoine, M., 1957c. Note preliminaire sur la geologie du Timor Portugais.
Garcia de Orto 5, 153163.
Gageonnet, R., Lemoine, M., 1958. Contribution a la connaissance de la geologie de la
province Portugaise de Timor. Estudos Ensaios Docum. Jta Invest. Ultramar 48, 1138.
Giani, L., 1971. The geology of the Belu District of Indonesian Timor. University of
London, unpublished M. Ph. Thesis.
Gradstein, F.M., Ludden, J., Adamson, A.C., Baumgartner, P.O., Beausillon, Thomas Bolmer,
T., Bown, P.R., Brereton, R., Bufer, R.T., Castillo, D., Compton, J., Dumoulin, J.A.,
Grifths, C.M., Haig, D., Heggie, D., Ishiwatari, A., Kaminski, M.A., Kodama, K., KopaskaMerkel, D.C., Marcoux, J.P., McMinn, A., Moran, M.J., Mutterlose, J., James, Q., Ogg, J.G.,
O'Neill, B., Plank, T., Riggins, M., Schott, M., Simmons, G., Thurow, J., 1990. Argo Abyssal
Plain/Exmouth Plateau: Proceedings of the Ocean Drilling Program, Initial Reports,
vol. 123. Ocean Drilling Program, College Station, Texas.
Gradstein, F., Ogg, J., Smith, A. (Eds.), 2004. A Geologic Time Scale. Cambridge University
Press, Cambridge.
Grady, A.E., 1975. A reinvestigation of thrusting in Portuguese Timor. Journal of the
Geological Society of Australia 22, 223227.
Grady, A.E., Berry, R.F., 1977. Some PalaeozoicMesozoic stratigraphicstructural
relationships in East Timor and their signicance in the tectonic of Timor. Journal
of the Geological Society of Australia 24, 203214.
Grunau, H.R., 1953. Geologie von Portugiesisch Ost-Timor. Eine kurze bersicht.
Eclogae Geologicae Helveticae 46, 2937.
Grunau, H.R., 1956. Zur geologie von Portugiesisch Ost-Timor. Mitteilungen Naturforschende Gesellschaft Bern 13, 1118.
Grunau, H.R., 1957a. Neue daten zur gelogie von Portugiesisch Ost-Timor. Eclogae
Geologicae Helveticae 50, 6998.
Grunau, H.R., 1957b. Geologia da parte oriental do Timor Portugues. Garcia de Orto 5,
727737.
Haig, D.W., McCartain, E., 2007. Carbonate pelagites in the Post-Gondwana succession
(CretaceousNeogene) of East Timor. Australian Journal of Earth Sciences 54,
875897.
Haig, D.W., McCartain, E., Barber, L., Backhouse, J., 2007. TriassicLower Jurassic
foraminiferal indices for Bahaman-type carbonate-bank limestones, Cablac Mountain Range, East Timor. Journal of Foraminiferal Research 37, 248264.
Haig, D.W., McCartain, E., Keep, M., Barber, L., 2008. Re-evaluation of the Cablac
Limestone at its type area, East Timor: revision of the Miocene stratigraphy of
Timor. Journal of Asian Earth Sciences 33, 366378.
Hall, R., 1997. Cenozoic plate tectonic reconstructions of SE Asia. In: Fraser, A.J., Matthews,
S.J., Murphy, R.W. (Eds.), Petroleum Geology of Southeast Asia: Geological Society
Special Publication, vol. 126, pp. 1123.

109

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, 353431.
Hamilton, W., 1979. Tectonics of the Indonesian region. United States Geological
Survey, p. 1078. Professional Paper.
Haq, B.U., von Rad, U., O'Connell, S., Bent, A., Blome, C.D., Borella, P.E., Boyd, R., Bralower,
T.J., Brenner, W.W., de Carlo, E.H., Dumont, T., Exon, N., Galbrun, B., Golovchenko, X.,
Grr, N., Ito, M., Lorenzo, J.M., Meyers, P.A., Moxon, I., O'Brien, D.K., Oda, M., Sarti,
M., Siesser, W.G., Snowdon, L.R., Tang, C., Wilkens, R.H., Williamson, P., Wonders, A.
A.H., 1990. Proceedings of the Ocean Drilling Program, Initial Reports, vol. 122.
Ocean Drilling Program, College Station, Texas.
Harris, R.A., 1989. Processes of allochthon emplacement with special reference to the
Brooks Range ophiolite, Alaska, and Timor, Indonesia. Unpublished PhD thesis,
University of London, 514 pp.
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,
373386.
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, 207231.
Harris, R.A., Long, T., 2000. The Timor ophiolite, Indonesia: model or myth? Geological
Society of America Special Paper, vol. 349, pp. 321330.
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 Banda Orogen, Indonesia. Tectonics 17, 458479.
Harris, R.A., Kaiser, J., Hurford, A., Carter, A., 2000. Thermal history of Australian passive
margin cover sequences accreted to Timor during Late Neogene arccontinent
collision, Indonesia. Journal of Asian Earth Sciences 18, 4769.
Harroweld, M., Keep, M., 2005. Tectonic modications of the Australian North West Shelf;
episodic rejuvenation of long-lived basin divisions. Basin Research 17, 225239.
Harroweld, M., Cunneen, J., Keep, M., Crowe, W., 2003. Early-stage orogenesis in the
Timor Sea region, NW Australia. Journal of the Geological Society of London 160,
9911001.
Harroweld, M., Holdgate, G.R., Wilson, C.J.L., McLoughlin, S., 2005. Tectonic
signicance of the Lambert graben, East Antarctica: reconstructing the Gondwanan
rift. Geology 33, 197200.
Hartono, H.M.S., 1990. Late Cenozoic tectonic development of the Southeast Asian
continental margin in the Banda Sea area. Tectonophysics 181, 267276.
Heap, A.D., Harris, P.T., 2008. Geomorphology of the Australian margin and adjacent
seaoor. Australian Journal of Earth Sciences 55, 555585.
Hinschberger, F., Malod, J.-A., Rhault, J.-P., Villeneuve, M., Royer, J.-Y., Burhanuddin, S.,
2005. Late Cenozoic geodynamic evolution of eastern Indonesia. Tectonophysics
404, 91118.
Hirschi, H., 1907. Zur geologie und geographie von Portugiesisch-Timor, Geologische
Mitteilungen aus dem Indo-Australischen Archipel. Nueues Jahrbuch fr Geologie
und Palontologie, Abhandlungen 24, 460474.
Hocking, R.M., Moors, H.T., Van De Graaff, J.E., 1987. Geology of the Carnarvon Basin,
Western Australia. Geological Survey of Western Australia Bulletin 133 289 pp.
Honthaas, C., Rehault, J.-P., Maury, R.C., Bellon, H., Hemond, C., Malod, J.-A., Cornee, J.-J.,
Villeneuve, M., Cotten, J., Burhanuddin, S., Guillou, H., Arnaud, N., 1998. A Neogene
back-arc origin for the Banda Sea basins: geochemical and geochronological
constraints for the Banda ridges (East Indonesia). Tectonophysics 298, 297317.
Huang, C.-Y., Yuan, P.B., Lin, C.-W., Wang, T.K., Chang, C.-P., 2000. Geodynamic
processes of Taiwan arccontinent collision and comparison with analogs in Timor,
Papua New Guinea, Urals and Corsica. Tectonophyics 325, 121.
Hunter, D.C., 1993; A stratigraphic and structural study of the Maubisse area, East
Timor, Indonesia. West Virginia University, unpublished MSc thesis, 107 pp.
Kaneko, Y., Maruyama, S., Kadarusman, A., Ota, T., Ishikawa, M., Tsujimori, T., Ishikawa,
A., Okamoto, K., 2007. On-going orogeny in the outer-arc of the TimorTanimbar
region, eastern Indonesia. Gondwana Research 11, 218233.
Karig, D.E., Barber, A.J., Charlton, T.R., Klemperer, S., Hussong, D.M., 1987. Nature and
distribution of deformation across the Banda ArcAustralia collision zone at Timor.
Geological Society of America Bulletin 98, 1832.
Keep, M., Clough, M.C., Langhi, L., 2002. Neogene structural and tectonic evolution of
the Timor Sea region, NW Australia. In: Keep, M., Moss, S.J. (Eds.), The Sedimentary
Basins of Western Australia: Proceedings of the Petroleum Exploration Society of
Australia, Perth, vol. 3, pp. 341354.
Keep, M., Longley, I., Jones, R., 2003. Sumba and its effect on Australia's north-western
margin. Geological Society of Australia Special Publication 22, 303312.
Keep, M., Beck, L., Bekkers, P., 2005. Complex modied thrust systems along the
southern margin of East Timor. Australian Petroleum Production and Exploration
Journal 45, 297310.
Keep, M., Barber, L., Haig, D., 2009. Deformation of the Cablac Mountain Range, East
Timor: an overthrust stack derived from an Australian continental terrace. Journal
of Asian Earth Sciences. doi:10.1016/j.jseaes.2009.02.001.
Kennett, J.P., Srinivasan, M.S., 1983. Neogene planktonic foraminifera, a phylogenetic
atlas. Hutchinson Ross Publishing Company, Stroudsburg, Pennsylvania. 265 pp.
Kenyon, C.S., 1974. Stratigraphy and sedimentology of the late Miocene to Quaternary
deposits of Timor. University of London,unpublished thesis.
Kristan-Tollman, von E., Barkham, S., Gruber, B., 1987. Ptschenschicten, Zlambachmergel (Hallsttter Obertrias) und Liaseckenmergel in Zentraltimor, nebst ihren
Faunenelementen. Mitteilungen der sterreichischen Geologischen Gesellschaft
80, 229285.
Krumbeck, L., 1921. Die Brachiopoden, Lamellibranchiaten und Gastropoden der Trias
von Timor. I. Stratigraphischer Teil. Palontologie von Timor 10, Abhandlungen 17,
1142.

Author's personal copy


110

M. Keep, D.W. Haig / Tectonophysics 483 (2010) 93111

Li, Z.X., Powell, C., 2001. An outline of the palaeogeographic evolution of the Australasian
region since the beginning of the Neoproterozoic. Earth Science Reviews 53, 237277
McA.
Lisboa, J.V., Carvalho, J.M.F., Oliveira, A., Carvalho, C., Grade, J., 2007. A preliminary case
study of potential ceramic raw materials in the Aileu area of Timor Leste. Journal of
Asian Earth Sciences 29, 593603.
Lister, G.S., Etheridge, M.A., Symonds, P.A., 1986. Detachment faulting and the evolution
of passive continental margins. Geology 14, 246250.
Longley, I.M., Buessenchuett, C., Clydesdale, et al., 2002. The North West Shelf a
Woodside perspective. In: Keep, M., Moss, S.J. (Eds.), The Sedimentary Basins of
Western Australia: Proceedings of the Petroleum Exploration Society of Australia
Symposium, Perth, vol. 3, pp. 2787.
Lourens, L., Hilgen, F., Shackleton, N.J., Laskar, J., Wilson, D., 2004. The Neogene Period.
In: Gradstein, F., Ogg, J., Smith, A. (Eds.), A Geologic Time Scale 2004. Cambridge
University Press, Cambridge, pp. 409440.
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,
481500.
Marks, P., 1961. The succession of nappes in the western Miomaffo area of the island of
Timor: a possible key to the structure of Timor. Proceedings of the Ninth Pacic
Science Congress, Geology and Geophysics, vol. 12, pp. 306310.
Martini, R., Zaninetti, L., Villenueve, M., Cornee, J.-J., Krystyn, L., Cirilli, S., De Wever, P.,
Dumitrica, P., Harsolumakso, A., 2000. Triassic pelagic deposits of Timor:
palaeogeographic and sea-level implications. Palaeogeography, Palaeoclimatology,
Palaeoecology 160, 123151.
McCaffrey, R., 1989. Seismological constaints and speculations on Banda Arc tectonics.
Netherlands Journal of Sea Research 24, 141152.
McCaffrey, R., Molnar, P., Roecker, S.W., Joyodiwiryo, Y.S., 1985. Microearthquake
seismicity and fault plane solutions related to arccontinent collision in the eastern
Sunda Arc, Indonesia. Journal of Geophysical Research 90, 45114528.
McCartain, E., Backhouse, J., Haig, D.W., Balme, B., Keep, M., 2006. Gondwana-related
Late Permian palynoora, foraminifers and lithofacies from the Wailuli Valley,
Timor Leste. Neues Jahrbuch fur Geologie und Palaontologie Monatshefte 240,
5380.
Metcalfe, I., 1996. Gondwanaland dispersion, Asian accretion and evolution of eastern
Tethys. Australian Journal of Earth Sciences 43, 605623.
Metcalfe, I., 2006. Palaeozoic and Mesozoic tectonic evolution and palaeogeography of
East Asian crustal fragments: the Korean Peninsula in context. Gondwana Research
9, 2446.
Molengraaf, G.A.F., 1912. De fatoe's van Timor. Geologische Mijnbouwkundig Genootschap voor Nederland en Kolonin, Geologische Sectie, Verslag, vol. 19121914,
pp. 117119.
Molengraaf, G.A.F., 1915. On recent crustal movements in the island of Timor and their
bearing on the geological history of the East-Indian archipelago. Koninklijke Akademie
van Wetenschappen, Amsterdam, Proceedings, vol. 15, pp. 224235.
Monteiro, F. Da Costa, 2003. The Triassic strata from East Timor: stratigraphy,
sedimentology and hydrocarbon potential. University of Auckland, unpublished
MSc thesis, 102 pp.
Mller, R.D., Mihut, D., Baldwin, S., 1998. A new kinematic model for the formation and
evolution of the west and northwest Australian margin. In: Purcell, P.G., Purcell, R.R.
(Eds.), The Sedimentary Basins of Western Australia: Proceedings of Petroleum
Exploration Society of Australia Symposium, Perth, vol. 2, pp. 5572.
Nolet, G., 2009. Slabs do not go gently. Science 324, 11521153.
Nugroho, H., Harris, R., Lestariya, A.W., Maruf, B., 2009. Plate boundary reorganization
in the active Banda Arccontinent collision: insights from new GPS measurements.
Tectonophysics. doi:10.1016/j.tecto.2009.01.026.
Obayashi, M., Yoshimitsu, J., Fukao, Y., 2009. Science 324, 11731175.
Petkovic, P., Collins, C.D.N., Finlayson, D.M., 2000. A crustal transect between Precambrian
Australia and the Timor Trough across the Vulcan Sub-basin. Tectonophysics 329,
2328.
Pigram, C.J., Panggabean, H., 1984. Rifting of the northern margin of the Australian
continent and the origin of some microcontinents in Eastern Indonesia. Tectonophysics 107, 331353.
Price, N.J., Audley-Charles, M.G., 1987. Tectonic collision processes after plate rupture.
Tectonophysics 140, 121129.
Rangin, C., Jolivet, L., Pubellier, M., Azema, J., Briais, A., Chotin, P., Fontaine, H., Huchon,
P., Maury, R., Muller, C., Rampnoux, J.P., Stephan, J.F., Tournon, J., 1990. A simple
model for the tectonic evolution of southeast Asia and Indonesia region for the past
43 m.y. Bulletin de la Societie Geologique de France 6, 889905.
Reed, D.L., 1985. Structure and stratigraphy of the Eastern Sunda forearc, Indonesia,
geologic consequences of arccontinent collision. PhD thesis, University of California,
Santa Cruz.
Reed, T.A., de Smet, M.E.M., Harahap, B.H., Sjapawi, A., 1996. Structural and depositional
history of East Timor. Proceedings of the Indonesian Petroleum Association 25, 297312.
Rey, S.S., Planke, S., Symonds, P.A., Faleide, J.I., 2008. Seismic volcanostratigraphy of the
Gascoyne margin, Western Australia. Journal of Volcanology and Geothermal
Research 172, 112131.
Richardson, A.N., Blundell, 1996. Continental collision in the Banda Arc. In: Hall, R.,
Blundell, D.J. (Eds.), Tectonic Evolution of Southeast Asia: Geological Society Special
Publication, vol. 106, pp. 4760.
Roniewicz, E., Stanley, G.D., Da Costa Monteiro, F., Grant-Mackie, J.A., 2005. Late Triassic
(Carnian) corals from Timor-Leste (East Timor): their identity, setting, and biogeography. Alcheringa 29, 287303.
Roosmawati, N., Harris, R., 2009. Surface uplift history of the incipient Banda arc
continent collision: geology and synorogenic foraminifera of Rote and Savu islands,
Indonesia. Tectonophysics. doi:10.1016/j.tecto.2009.04.009.

Rose, G., 1994. Late Triassic and Early Jurassic radiolarians from Timor, Eastern Indonesia.
PhD Thesis, University of London (unpublished).
Rosidi, H.M.D., Suwitodirdjo, K., Tjokrosapoetro, S., 1979. Geological map of the Kupang
Atambua quadrangle, Timor. Geological Research and Development Centre,
Bandung.
Rothpletz, A., 1891. The Permian, Triassic and Jurassic formations in the East Indian
Archipelago (Timor and Rotti). American Naturalist 25, 959962.
Rothpletz, A., 1892. Die Perm-, Trias und Jura-Formation auf Timor und Rotti im
Indischen Archipel. Palaeontographica 39, 57106.
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, 453479.
Sandwell, D.T., SMITH, W.H.F., 1997. Marine gravity anomaly from Geosat and ERS-1
satellite altimetry. Journal of Geophysical Research 102, 1003910054.
Sani, K., Jacobsen, M.L., Sigit, R., 1995. The thin-skinned thrust structures of Timor.
Proceedings of the Indonesian Petroleum Association 24, 277293.
Sartono, S., 1975. The age of the Kekneno Formation in Timor, Indonesia. Geologi
Indonesia 2, 2937.
Sashida, K., Adachi, S., Ueno, K., Munasri, 1996. Late Triassic radiolarians from Nefokoko,
west Timor, Indonesia, pp. 225234. Prof. H. Igo Commemorative Volume.
Sashida, K., Kamata, Y., Adachi, S., Munasri, 1999. Middle Triassic radiolarians from
West Timor, Indonesia. Journal of Paleontology 73, 765786.
Sawyer, R.K., Sani, K., Brown, S., 1993. The stratigraphy and sedimentology of West
Timor, Indonesia. Proceedings of the Indonesian Petroleum Association, 22nd
Annual convention, pp. 533574.
Silver, E.A., Gill, J.B., Schwartz, D., Prasetyo, 1985. Evidence for a submerged and displaced
continental borderland, north Banda Sea, Indonesia. Geology 13, 687691.
Simons, A.L., 1940. Geological investigations in N.E. Netherlands, Timor. In: Brouwer, H.A.
(Ed.), Geological Expedition of the University of Amsterdam to the Lesser Sunda
Islands in the South Eastern Part of the Netherlands East Indies 1937, vol. 1. N.V.
Noord-Hollandsche Uitgevers Maatschappij, Amsterdam, pp. 110213.
Sinha, D.K., Singh, A.K., 2008. Late Neogene planktic foraminiferal biochronology of the
ODP Site 763A, Exmouth Plateau, Southeast Indian Ocean. Journal of Foraminiferal
Research 38, 251270.
Snyder, D.B., Milsom, J., Prasetyo, H., 1996. Geophysical evidence for local indentor
tectonics in the Banda Arc east of Timor. In: Hall, R., Blundell, D. (Eds.), Tectonic
Evolution of Southeast Asia: Geological Society Special Publication, vol. 106,
pp. 6173.
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, 112.
Sopaheluwakan, J., Helmers, H., Tjokrosapoetro, S., Surya Nila, E., 1989. Medium
pressure metamorphism with inverted thermal gradient associated with
ophiolite nappe emplacement in Timor. Netherlands Journal of Sea Research
24, 333343.
Standley, C.E., Harris, R., 2009. Tectonic evolution of forearc nappes of the active Banda
Arccontinent collision: origin, age, metamorphic history and structure of the
Lolotoi Complex, East Timor. Tectonophysics. doi:10.1016/j.tecto.2009.01.034.
Tappenbeck, D., 1940. Geologie des Mollogebirges und einiger benachbarter gebiete
(Niederlndisch Timor). In: Brouwer, H.A. (Ed.), Geological Expedition of the
University of Amsterdam to the Lesser Sunda Islands in the South Eastern Part of the
Netherlands East Indies 1937, vol. 1. N.V. Noord-Hollandsche Uitgevers Maatschappij, Amsterdam, pp. 1105.
Teixeira, C., 1952. Notas sobre a geologia e a tectonica de Timor. Revta. Esc. Sup. Solon.,
Lisb. 3, 85154.
Tobing, S.L., 1989. The geology of East Timor. University of London M. Phil thesis,
unpublished, 129 pp.
Twiss, R.J., Moores, E.M., 2007. Structural Geology, 2nd Edition. W. H. Freeman and Co,
Press, New York. 736 pp.
van Bemmelen, R.W., 1949. The Geology of Indonesia. Government Publishing Ofce,
The Hague.
van der Werff, W., 1995. Cenozoic evolution of the Savu Basin, Indonesia: forearc
basin response to arccontinent collision. Marine and Petroleum Geology 12,
247262.
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, 363374.
van Marle, L.J., 1991. Late Cenozoic palaeobathymetry and geohistory analysis
of Central West Timor, eastern Indonesia. Marine and Petroleum Geology 8,
2234.
van West, F.P., 1941. Geological investigations in the Miomaffo region (Netherlands
Timor). In: Brouwer, H.A. (Ed.), Geological Expedition of the University of
Amsterdam to the Lesser Sunda Islands in the South Eastern Part of the Netherlands
East Indies 1937, Volume III. N.V. Noord-Hollandsche Uitgevers Maatschappij,
Amsterdam, pp. 1131.
Villeneuve, M., Harsolumkso, A.H., Cornee, J.J., Bellon, H., 1999. Structure of West Timor
(Indonesia) along a northsouth cross section. Geologie Mediterraneenne 26,
127142.
Villeneuve, M., Corne, J.-J., Harsolumakso, A., Martini, R., Zaninetti, L., 2005. Rvision
stratigraphique de I'Lle de Timor (Indonsie orientale). Ecolgae Geologicae Helvetiae
98, 297310.
Wanner, J., 1913. Geologie von Westtimor. Geologische Rundschau 4, 136150.
Wensink, H., 1994. Paleomagnetism of rocks from Sumba: tectonic implications since
the late Cretaceous. Journal of Southeast Asian Earth Sciences 9, 5165.
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, 5771.

Author's personal copy


M. Keep, D.W. Haig / Tectonophysics 483 (2010) 93111
Wensink, H., Hartosukohardjo, S., 1990. Palaeomagnetism of younger volcanic rocks
from Western Timor, Indonesia. Earth and Planetary Science Letters 100, 94107.
Wensink, H., van Bergen, M.J., 1995. The tectonic emplacement of Sumba in the SunaBanda Arc: paleomagnetic and geochemical evidence from the Miocene Jawila
volcanics. Tectonophysics 250, 1530.

111

Wittouck, S.F., 1937. Exploration of Portuguese Timor. Report of Allied Mining Corp. To
Asia Investment Co. Ltd., Amsterdam (Kolff).
Woodside, J.M., Jongsma, D., Thommeret, M., Stang van Hees, G.L., Puntodewo, 1989.
Gravity and magnetic eld measurements in the eastern Banda Sea. Netherlands
Journal of Sea Research 24, 185203.

Vous aimerez peut-être aussi