Sie sind auf Seite 1von 44

Five anomalous structural aspects of rift basins in Thailand and their impact on

petroleum systems
By
Chris K. Morley
PTTEP, 27th Floor, ENCO Building, Soi 11, Vibhavadi-Rangsit Road, Chatuchak, Bangkok, Thailand,
10900
chrissmorley@gmail.com
Abstract
The intra-cratonic, supra-subduction zone setting of the Thailand rift basins caused the rifts to evolve in
several ways differently from other intra-cratonic rifts. Key differences include: widespread occurrence of
low-angle normal faults; basin inversion alternating with rifting; diachronous initiation and cessation of
rifting; rapid post-rift subsidence and extensive, low-displacement post-rift faults. These characteristics are
related to hot, weak continental lithosphere, rapid evolution of the plate boundaries and stresses during the
Cenozoic, and the history of subduction and accretion. Low-angle normal faults impacted the Sirikit Field by
controlling the location of fluvio-deltaic reservoirs interfingering lacustrine shales (source and seal). Basin
inversion alternating with syn-rift section is detrimental to prospectivity of the southern half of the Phitsanulok Basin by destroying syn-rift structure, and halting hydrocarbon maturation early in the basin history.
The diachronous timing of basin development resulted in large, gas-dominated hydrocarbon accumulations
in the eastern Gulf of Thailand basins, but was detrimental for younger basins that lack the post-rift section,
and the greater variety of petroleum systems and traps that post-rift basins offer. The extensive low displacement-length ratio post-rift faults form significant hydrocarbon traps in the Pattani and North Malay basins.
They represent a structural style not usually found in post-rift basins.
Introduction
The Cenozoic rift system that runs through central Thailand is about 1,900 km long, and up to 300
km wide. The northern part of the rift system lies 800 km south of the Eastern Himalayan Syntaxis, while
the westernmost part of the rift system lies 900 km east of the Andaman trench (Fig. 1). West of the rift
basins a back-arc rift system is present in the Andaman Sea area, this rift system is located at most 600 km
from the trench (Figs. 1 and 2). The continental core of Southeast Asia is called Sundaland (e.g. Hall and
Morley, 2004; Fig. 1). Thailand rift basins form a failed rift system in the central part of Sundaland. Although the basins are remotely associated with the Himalayan collisional setting, and the Andaman-Sumatra
subduction zone, they lie too far from these areas to be called either collisional or back-arc basins, but their

development is clearly influenced by these settings (e.g. Hall and Morley, 2004).
As outlined above the rift basins of Thailand have developed in a tectonic setting very different
from typical continental rifts in old cratonic areas such as the East African Rift System, Central African Rift
System, and the North Sea. Yet all these basins can be classed as failed rifts, or intracratonic rifts. As a way
of highlighting the differences within this class of rift, this paper investigates five structural aspects of the
rift basins in Thailand that are anomalous with respect to typical failed rifts such as the ones listed above,
describes how these aspects have influenced hydrocarbon prospectivity, and finally discusses the tectonic
characteristics of SE Asia that have influenced these anomalous features.

The five aspects considered in this paper are: 1) highly diachronous timing of rift basin cessation

along the system, 2) Low-angle normal faults, 3) inversion alternating with extension, 4) very rapid post-rift
subsidence, and 5) extensive extensional faulting in the post-rift section.
Geological background
Both rifting and strike-slip deformation have extensively affected onshore and offshore Thailand during the Cenozoic (Polachan et al., 1991; Lacassin et al., 1997; Morley, 2004; Morley et al., 2011). The onset
of rift basin formation ranges in age from Late Eocene (e.g. Krabi Basin) to Late Miocene, (e.g. Phitsanulok
Basin), (Morley and Racey, 2011). The rift basins are at their broadest in the Gulf of Thailand, and extend
upwards through central Thailand to the northernmost part of the country (Fig. 3). The rift basins of the
Gulf of Thailand to northern Thailand are dominantly or entirely filled by continental deposits (Fig. 4). It is
only in the post-rift phase that marine incursions become a feature of the Gulf of Thailand basins (Morley
and Racey, 2011). In contrast the Mergui and Andaman syn-rift basins of the Andaman Sea region (Fig. 3)
are dominantly filled by marine and fluvio-deltatic deposits. The Mergui and Andaman Basins are not discussed in detail in this paper because they contain no commercial hydrocarbon discoveries at present.
The nature of the interactions between strike-slip faulting and rifting in Thailand is controversial.
Clearly there are examples where strike-slip faults are linked to rift basins, in some cases these relationships
can be interpreted as transfer faults (e.g. Fang Basin, Morley, 2007), but there are also clear examples where
the basins lie at releasing bends associated with strike-slip faults (e.g. Mae Sot Basin, Mae Ping Fault zone,
and Pracham, Sisawat and Huai Malai basins along the Three Pagodas Faults, Morley and Racey, 2011,
Morley et al., 2011). Some workers have sought to explain all the rift basins in a strike-slip context (e.g.
Tapponnier et al., 1986; Polachan et al., 1991). However, it is the opinion of the author that the evolution of
the basins is more complex, and that most of the Cenozoic basins opened under an extensional stress regime
with the strike-slip faults either activating at different times from extension, and/or that there is a spatial

transition from areas of strike-slip to extensional dominated stress regimes (e.g. Morley, 2002b, 2007; Morley et al., 2011).

The most prolific hydrocarbon province lies in the eastern Gulf of Thailand, where the Pattani Basin

and North Malay basin produce predominantly gas with some liquids, primarily from structural and structural-stratigraphic traps associated with conjugate fault sets in the post-rift section (e.g. Racey, 2011). Total
produced gas and remaining proven plus probable reserves of about 30 TCF gas are estimated for the eastern
Gulf of Thailand, together with about 360 MMbbl oil and 723 MMbbl condensate (Racey, 2011). A small
number of scattered oil fields are present in the western Gulf of Thailand. Oil production from the western
Gulf may in the near future attain a peak of about 100,000 bbl/day from the Kra (Manora Field, ~20-29
MMbbl, recoverable reserves) Western (Bualuang oil field, recoverable reserves ~20 MMbbl), Songkhla-Nakhon (recoverable reserves estimated between 10-28 MMbbl), and Chumphon basins. Onshore, small fields
have been found in the Fang, Suphan Buri and Phetchabun basins. Sirkit field is the largest oil field in all of
Thailand and lies within the Phitsanulok Basin (Flint et al, 1988; Morley et al., 2007a). Total output from
the field is about 200 MMbbl oil, with perhaps a further 20% yet to be produced.

The general structural style of the rifts comprises typical arrangements of half grabens found in rifts,

which change polarity and give rise to a variety of transfer zones (Morley et al., 1990). Overall the rifts trend
N-S, but the influence of pre-existing fabrics causes NE-SW and NW-SE and even ENE-WSW trends to be
activated in places (Morley et al., 2004).
Diachronous termination of rifting
Rift initiation and termination prior to the development of oceanic crust is commonly related to
either an aulacogene setting (failed arm of a triple junction; Burke, 1977; Fig. 5), or the abandoned propagating tip of a rift system (Morley 2002a). The best examples of propagating rifts are along the successful
rifts that have gone on to form passive margins, for example the Jurassic to Cretaceous break-up along a
9000 km long segment of the African margin. Aulacogenes lie at a high angle to the passive margin rifts,
and are related in timing to the evolution of the adjacent passive margin segment. In the simplest situations
these failed rifts tend to develop along their length over a short period of time and propagate together from a
number of initial seed points or propagate inland with time (Genik, 1994; Koehn et al., 2008; Roberts et al.,
2012; Fig. 5), and terminate activity once seafloor spreading begins on the adjacent rift margin, unless an active component to the driving forces remains (e.g. East African Rift, Nyblade, 2002). However, once created
the rifts are zones of weakness that may be subsequently reactivated multiple times.
A good example of the basic pattern outlined above are the Jurassic syn-rift sub-basins of the North

Sea. These basins show a diachronous onset of Late Jurassic extension with rifting propagating towards the
centre of a pre-rift dome with time (Rattey and Hayward, 1993). The termination of this short (< 15 my) period of extension across almost the entire rift system is marked by the base Cretaceous or late Cimmerian
unconformity (e.g. Kyrkjebo et al., 2004). This termination is attributed to the onset of seafloor spreading in
the Atlantic (e.g. Rattey and Hayward, 1993). Periods of failed rift activity related to abandoned propagating rift tips are represented by the Gulf of Suez (~15 Ma duration, Bosworth et al., 2005) and the Reconcavo
Basin (~18 Ma duration, Magnevita et al., 2005).
The Central African Rift is 3,800 km long, traverses the entire length of central Africa, and exhibits a
remarkably uniform Early Cretaceous onset of rifting in most basins, and termination of initial rifting around
the mid to late Albian (Genik, 1993; Keller et al., 2006). However, subsequent extensional reactivation of
the rifts continued into the Palaeogene, long after mid-Cretaceous seafloor spreading was initiated in the Atlantic (Genik, 1994; Keller et al., 2006). Fairhead et al. (2013) show how three distinct phases of extension
in the Central African Rift system (particularly the Muglad Basin, Sudan) can be correlated with azimuthal
changes in Atlantic ocean fracture zones. The Sirte Basin is another example of pulsed basin reactivation
(Abdulbaset et al., 2008).
The East African Rift system exhibits complex timing of basins in part related to propagation of the
rift system to the south to southwest with time, and interactions with mantle plumes (Nyblade, 2002). The
Eastern Branch in the Turkana area shows Eocene-Oligocene volcanism followed by Oligocene basin development (Morley et al., 1999). Oligocene basin development is consistent with the onset of extension in the
Gulf of Aden (Bosworth et al., 2005). Deformation has propagated to the south in Kenya and Tanzania with
time (Morley, 1999; Le Gall et al., 2004). Although Oligocene syn-rift sediments have been identified for
the Lake Rukwa segment in the Western Branch (Roberts et al., 2012), much of the 2000 km long Western
Branch development is Neogene in age (Ebinger, 1989). There are also newly developing rift segments of
Holocene age to the SW of the Western Branch, such as the Okavango Rift Zone (Kinabo et al., 2008).

The rift basins of Thailand show considerable diachroneity in timing both regarding the onset and

termination of rifting (in other failed rifts it tends to be one but not the other), in general showing a younging
of both rift onset and termination to the north with time (Fig. 4). There is also an east to west shift in activity within the Gulf of Thailand. As reviewed in Morley and Racey (2011) in the Gulf of Thailand the oldest
rifts (Late Eocene-Oligocene) occur in the eastern and southern portion of the Gulf. Rifting ceased around
the Oligocene-Miocene boundary in the eastern Gulf, while in the western Gulf the story is more diverse.
Some basins in the southwest of the Gulf show both Late Eocene and Early Miocene phases of extension.
While in the NW part of the Gulf the rift basins are predominantly of Late Oligocene-Early Miocene age.

Onshore in Central Thailand and in NW Thailand to the Chiang Mai and Fang basins, rifting lasted from
Late Oligocene to Late Miocene. In eastern northern Thailand some rift basins appear to only be of Middle
Miocene or Middle to Late Miocene age (Morley et al., 2001). The northwards younging of rifting is also
observed in the Mergui Basin (Srisuriyon and Morley, in review). The difference in age of the onset of rifting from south to north is moderate, perhaps about 10 my, while the difference in the age of rift termination
is around 15-20 my. The result is that some basins have been accumulating sediment in thick post-rift basins,
while contemporaneously other basins have continued to accumulate sediments in active half grabens.

In a review of hydrocarbon resources in different rift basin and post-rift basin associations Morley

(1999) concluded that simple rifts (i.e. failed rifts without a post-rift basin) tended to have small resources
compared with rifts capped either by a sag basin (failed rift) or passive margin (Fig. 6). The absence of the
capping post-rift basin meant that there were problems with sealing potential reservoir units in large traps,
and that maturation of the source rock was likely to be limited to the deepest parts of the basin. The post-rift
basin can also be the site of completely different plays, and adds to the potential diversity of the petroleum
systems. The Reconcavo Basin, Brazil, is the most prolific example of hydrocarbons in a simple rift. From
the start of exploration in 1937 to December 2001 5,626 wells were drilled in the basin resulting in cumulative production of 1.4 billion bbl oil and 1.9 TCF gas. By comparison the North Sea comprises two rift
systems (Triassic and Late Jurassic) capped by post-rift basins. The largest field in the Cretaceous-Cenozoic
post rift section is Forties Field (~5 billion bbl oil in place), while one of the largest syn-rift fields, Gullfaks,
has produced more oil (~2.4 billion barrels) than the entire Reconcavo Basin.
The discoveries in Lake Albert in the past 7 years have added another example of a highly successful simple rift play, with about 1.2 billion barrels of recoverable oil (http://www.tullowoil.com/index.
asp?pageid=282). Contributing factors to the success of the Reconcavo Basin are the unusually high quality of lacustrine turbidite sandstone reservoirs, and the presence of good pre-rift reservoir sands sealed by
overlying lacustrine shales. Very high geothermal gradients are a significant factor in maturing the young
sediment in Lake Albert.

The fundamental differences in hydrocarbon plays between simple rift basins and rifts with sag

basins discussed above is clearly seen in the Thailand rift basins. The rift basins that were active during the
Miocene have thin, poorly developed to virtually absent post-rift basin section, and they can essentially be
classed as simple syn-rift basins. The basins of the eastern Gulf have a very well developed Neogene postrift section up to 6 km thick that is the primary hydrocarbon exploration target. As mentioned in the geological background section the eastern gulf (syn-rift plus post-rift) contains the bulk of the hydrocarbon reserves
of the country and is gas-prone, while the western gulf and onshore rift basins (simple rift) contains relative-

ly small oil fields, with Sirikit Field being the one notable exception (Racey, 2011).

While the generalization regarding the syn-rift vs post-rift-dominated plays works well, there are

anomalous distributions of fields within these two classes of basin that require explanation. For example, despite the general small field size (up to 10s mmbbo) in the simple rifts, there is the Sirkit Field in this class
of rift, which has produced over 200 mmbbo to date. The reason for the occurrence of the Sirikit field is
discussed in the low-angle normal faults section. In the eastern Gulf there are two main trends of basins, the
gas-rich Pattani-North Malay basin trend, and the Khmer Basin, which at present has only small, marginally
economic discoveries. Two anomalous features of the Pattani and North Malay basins are the great depth
and rapid subsidence exhibited by the post-rift basins, and the extensive swarms of low-displacement normal faults that set up the trapping configurations (Rigo de Rhigi et al., 2002; Kornsawan and Morley, 2002).
These aspects are discussed in subsequent sections.
Low-angle normal faults

Typical failed continental rifts are considered to be composed predominantly of normal faults initi-

ated at a high-angle (e.g. Buck, 1991). This appears to be the case for the majority of the major boundary
faults in the East African Rift (e.g. Morley, 1989), the Gulf of Suez (Jackson et al., 1988) and the North Sea
(Fossen et al., 2000). Low-angle faults in these settings tend form either the detachment zones flattening out
deep within the crust, or they are the result of rotation of initially high-angle faults, a few may also result
from reactivation of low-angle pre-existing orogenic structures (e.g. Fossen et al., 2000, Lyngsie, 2007). The
occurrence of high-extension, metamorphic core-complex-type low-angle normal faults is generally explained either by atypical crustal conditions (such as the hot crust or over-thickened crust of orogenic belts),
together with rotation of initially high-angle faults (e.g. Wernicke and Axen, 1988; see review in Whitney et
al., 2013).
Six low-angle normal faults (LANFs) bound half-graben basins in the Gulf of Thailand-onshore rift
basin trend, two bound crystalline basement in the footwall block and are seen in outcrop adjacent to the
Gulf of Thailand (Rayong, and Khanom), while four occur in the Mergui Basin. Morley (2009) described the
occurrence of LANFs in Thailand and noted that all of them dip eastwards between about 20 and 30. What
differentiates these basins from typical failed rifts is: a) the presence in some areas of metamorphic core
complex associations, and b) at least some of the normal faults were initiated at a low-angle and are not the
result of later rotation (Morley, 2009; Fig. 7).
The widely accepted models of LANF development discussed by Buck (1991) and Whitney et al.
(2013) require the crust to be warm or hot, and possibly over-thickened to generate low-angle normal faults

(Fig. 8). Whitney et al. (2013) view all the origins of LANFs as variations on rotation of high-angle faults to
a lower angle (Fig. 8). It is certainly probable hot, thickened crust existed in Thailand during the Cenozoic
as a consequence of an oblique Andean-type margin setting during the Late Cretaceous and Palaeogene (e.g.
Morley, 2004, 2012; Searle et al., 2007; Palin et al., 2013), and that the rift basins are in part related to collapse of this thickened crust (Morley et al., 2011; Morley, 2012). But that does not have to mean that all the
low-angle normal faults originated as rotated high-angle faults following the models reviewed by Whitney et
al. (2013; Fig. 8). A key conclusion by Morley (2009) was that at least in some cases (Doi Inthanon-Chiang
Mai Basin, Phitsanulok Basin, Suphan Buri Basin) the LANFs originated as low-angle faults and were not
rotated higher-angle faults. The development of the footwall areas, and relationships of the faults to the
sedimentary basins simply did not fit a rotating high-angle normal fault model (Morley, 2009; Fig. 7). The
cause of the low-angle faults is probably that they followed weak, low-angle, east-dipping compressional
fabrics generated during the Triassic-Early Jurassic Indosinian orogeny (Morley, 2009). The LANFs do not
represent a single stage in the rift development, the Rayong and Khanom areas may represent early collapse
of the crust at the onset of extension during the Eocene (Morley et al., 2011), and are the best candidates for
rotated, initially high-angle normal faults. Other basin-bounding faults were not active until the Late Oligocene or Early Miocene and are the ones that initiated at a low-angle (e.g. Chiang Mai Basin, Phitsanulok
Basin, Morley, 2009).

The impact of LANFs on hydrocarbon prospectivity is two-fold: 1) as a sediment source area during

unroofing (e.g. Chiang Mai Basin), and 2) the influence on depositional systems (e.g. Phitsanulok Basin) and
particularly the source/seal-reservoir relationships found in the Sirikit Field. Examples are discussed below.
Footwall denudation
The Chiang Mai basin is a Late Oligocene-Miocene rift basin about 125 km long, and 25 km wide
(Fig. 3). It is bounded on its western margin by a large, low-angle normal fault system (Dunning et al., 1995;
Rhodes et al., 2000). Mid-crustal level ortho- and para-gneisses are exposed in the footwall of fault LANF
(Dunning et al., 1995; Macdonald et al., 1993, 2010; Fig. 9). Mica Ar-Ar and apatite fission track ages indicate that unroofing of this metamorphic core complex occurred between about 21 and 14 Ma (Upton, 1999;
Morley, 2009). Making some simple assumptions about the depth to the top of the unroofed amphibolitic
gneisses at the start of extension (10 km depth), and the amount of extension (up to 35 km, Morley, 2009),
the total volume of material eroded or translated eastwards to achieve the unroofing is about 15,000 km3, of
which the volume of eroded crust is estimated at around 7,000 km3. The cooling ages indicate this material
was removed over a period of about 11-13 my. The total volume of sediments filling the Chiang Mai basin

estimated from seismic reflection data is about 5,000 km3, and perhaps represents deposition over a 15-20
million year period. Consequently the amount of material eroded from the Doi Inthanon/Doi Suthep area
far exceeded the capacity of the Chiang Mai Basin during the Early Miocene. Curiously, Miocene alluvial
fan deposits on the west-side of the basin, with one exception, do not contain clasts of gneiss from the Doi
Inthanon source area (Rhodes et al., 2005; B. Rhodes pers. comm., 2012). This suggests that while the basin
was being transported east the gneisses were not widely exposed, and that by the time the gneisses were
exposed and eroded the basin was lacking accommodation space. The Chiang Mai Basin LANF created a local sediment source area that generated a volume of sediment that far exceeded the capacity of the adjacent
Chiang Mai Basin. Consequently a pulse of sediment coming from the Doi Inthanon/Doi Suthep area during
the Early Miocene would have worked its way into the rift basins further south. Provenance studies to investigate whether reservoir sands in basins further south can be linked with the unroofing of the Doi Inthanon
area have yet to be conducted.

Footwall uplift associated with high-angle extensional fault systems will typically be a much smaller

source of sediment supply than the Doi Inthanon/Doi Suthep example. The footwall is clearly the source
area for many boundary fault margin fan systems in rifts, but such systems are typically confined to a few
kilometres width from the boundary fault. These fans do not represent a high volume of sediment input
compared with drainage systems that enter from the axial and flexural margins. The East African Rift clearly
shows a strong link between faulting and footwall topography (e.g. Hendrie et al., 1994; Kusznir et al., 1995;
Sachau and Koehn, 2010), although uplift at the largest scale is associated with mantle upwelling effects
(Ebinger et al., 1989; see reviews in Pik, 2011 and Moucha and Forte, 2011). Flexural cantilever modeling
of the rifting also highlights the potential for several kilometres of footwall section to be removed by erosion
during extension (e.g. Hendrie et al. 1994; Kusznir et al., 1995;). Yet, a regional apatite fission track study in
Kenya identified several episodes of denudation and cooling during Cretaceous-Palaeogene times, but relatively little basement denudation during the Miocene-Recent period of rifting (Foster and Gleadow, 1993,
1996). Only using the (U-Th)/He technique, which is sensitive to smaller amounts of denudation and cooling
than apatite-fission track, has a younger cooling event during the last 5-10 Ma been identified (Spiegel et al.,
2007). Hence, the denudation associated with Doi Inthanon is considered anomalous compared with typical
continental rifts.
Sirikit Field
The Phitsanulok Basin (Figs. 3 and 10) in Central Thailand is an example of a simple rift. Many of
the numerous fields in the basin are typical simple rift plays, with relatively low-displacement faults setting

up small traps, where only a few of the numerous potential sandstone reservoirs are laterally sealed and trap
hydrocarbons. One field, however, stands out as being unusually large. Sirkit Field is a tilted fault block trap
(Fig. 11), apart from its setting in a continental basin, in terms of structural style and size (~14 km2), it is a
good match for a typical North Sea-type tilted fault block trap (Morley et al., 2007b). However, typically
in the North Sea the Early Cretaceous post-rift section is an important element for the top seal. About 200
million bbl oil have been produced from the Sirkit Field, and the recovery factor is relatively low at about
16%. The Sirkit Field represents the perfect storm of simple rift traps in a continental rift depositional setting. The key to its size is that an axial fluivio-deltaic system prograded from the north over the top of the
tilted fault block (Fig. 12). Successive transgressions and regressions resulted in a sequence of alternating
sands (Lan Krabu Formation) and shales (Chum Saeng Formation), that juxtaposed source/seal rocks and
reservoir rocks over a structural high. Growth of the structural high was episodic, and insufficient to cause
complete erosion of section at the fault block crest (Morley et al., 2007b). The shales were sufficiently thick
to also be smeared into the main fault zone and hence to play an important role as both lateral and top seals.
The unusual feature of the basin is that the main hydrocarbon kitchen area, and main depocentre (Sukhothai
Depression) lies north of the Sirikit Field. Hence, it would normally be expected that high subsidence along
the bounding fault would prevent the southward axial progradation of the fluvio-deltaic system from reaching Sirikit Field. What makes this setting anomalous is the presence of the very low-angle Western Boundary Fault, which dips at about 25 E, has a maximum heave in excess of 12 km, and lies on the western side
of the Sukhothai Depression (Morley et al., 2007a; Morley, 2009).
Since the amount of (vertical) throw on a fault is related to the amount of extension (heave) and fault
dip angle, assuming similar strain rates high-angle faults will create more accommodation space than lowangle faults for a particular time-interval. Hence, the potential for high-angle faults outstripping sedimentation rates, and creating sediment starved basins, and locations for large deep lakes, is considerably greater
than for low-angle normal faults (Morley, 1989). Despite 12 km heave on the Western Boundary Fault, the
low throw:heave ratio meant that the hangingwall depocentre was at least frequently and temporarily unable
to keep up with sedimentation rates, hence the fluvio-deltaic sediments prograded over the entire depocentre
(Fig. 12). After an initial pulse of lacustrine sedimentation, the Sukhothai Depression was subject to predominantly fluvio-deltaic deposition, and lacustrine conditions remained south of the Sukhothai Depression,
where smaller, but higher-angle faults were able to sustain lacustrine conditions (Fig. 12). The Sirkit Field
tilted fault block lay in the perfect location for this transition. Had the Sukhothai Depression been bounded
by a large high-angle normal fault it is likely the fluvio-deltaic deposits would have remained confined to the
Sukhothai Depression, and never reached the Sirikit Field area depriving it of good reservoir rocks.

Basin Inversion
It is generally recognized within the petroleum industry that locally inverted rifts tend to enhance hydrocarbon prospectivity, while regional inversion tends to degrade it (Macgregor, 1995). In SE Asia, two of
the largest hydrocarbon provinces: the Central Basin, Sumatra, and the Malay Basin, offshore East Malaysia,
have the great majority of their reserves within inversion-related traps in the post-rift section (Macgregor,
1995). The timing of inversion when it affects rift and post-rift basins is usually either shortly after the end
of rifting, or at any time during or after the post-rift subsidence phase depending upon how the tectonic conditions of the area have evolved. What is atypical is to find examples of rifting alternating with inversion, yet
this is the case for some rift basins in Central and Northern Thailand.

The distribution of inversion in rift basins in Thailand is highly variable. As mentioned above, the

Malay Basin is strongly inverted, but passing northwards into the North Malay Basin and Gulf of Thailand
inversion is much less important. There are, never-the-less, examples of inversion around the OligoceneMiocene boundary (the end of the early rift phase in the gulf) in a few basins (Chumphon, Khmer Basin,
Mergui Basin; Morley et al., 2011). In the NW gulf, minor inversion occurs at the end of rifting during the
Middle Miocene. While in a number of basins, particularly onshore, inversion occurs at the end of rifting
around the Miocene-Pliocene boundary. Only in the Fang Basin (Pliocene inversion) in northern Thailand
have any of these events provided traps for hydrocarbons (Racey, 2011). Due to the diachronous nature of
the rifting events discussed in the section Diachronous Termination of Rifting, the Middle Miocene and
Pliocene inversion structures can be classed as the end of rifting-type structures.

The more unusual inversion structures occur in Central and Northern Thailand. Morley et al. (2000)

noted the possibility of 4-5 inversion events in the Late Oligocene-Miocene Li Basin based on outcrop
observations. Such inversion events have impacted the Phitsanulok Basin, which is the most prolific onshore
hydrocarbon basin. The northern half of the basin contains the Sukhothai Depression and Sirkit Oil Field
(Fig. 10). Inversion structures in the northern half of the basin occurred at the end of rifting, at the MiocenePliocene boundary (Morley et al., 2007a, 2011). The southern half of the basin has no hydrocarbon discoveries. The basin can be sub-divided into the Lahan, and Nong Bua and Dong Chat sub-basins (Fig. 10). The
Lahan Graben comprises Oligocene-Early Miocene section. Subsidence in the basin ceased after that time,
which is different from the Sukhothai Depression to the north and the Nong Bua sub-basin to the East, both
of which continued to subside within fault-bounded half graben during the Middle and Late Miocene. The
cessation of subsidence in the Lahan Graben is marked by strong inversion in a transfer zone region between
the Lahan Graben and the Nong Bua sub-basin (Fig. 10). Two phases of inversion affect the transfer zone

region (Fig. 13). The first inversion event created widespread folding, and erosion, which is marked by an
angular unconformity. The second inversion event is more subtle, and is seen on one fault, where the brown
unit is thinner in the hangingwall than in the footwall (Fig. 13B, C). One change further north, potentially
related to stress changes associated with inversion, is that the extensional fault pattern is re-organized from
more N-S trending faults to NE-SW trending faults about the same time as the inversion (Morley et al.,
2007a).
Inversion of the syn-rift section has not produced any large closed structures, and effectively marks
the end of significant basin subsidence, and the termination of hydrocarbon maturation in the basin. Probably the single largest contributing factor to the absence of any hydrocarbon discoveries in the southern half
of the Phitsanulok basin is the Middle Miocene inversion.

Extension during the Middle and Late Miocene was episodic in Central and Northern Thailand.

Most of the rift basins are not very deep, and the basin fill may amount to a few hundred metres (e.g. Mae
Moh and Li basins), or in the deeper basins (Chiang Mai, Lampang, Phrae) the stratigraphy is just not well
dated (Morley et al. 2001). Only in the Sukhothai Depression is the Middle-Late Miocene stratigraphy well
known (in terms of numerous well penetrations, and seismic coverage). This time period comprises the
uppermost part of the Pratu Tao Formation, the Yom and Ping formations (Flint et al. 1988; Morley et al.,
2007a). The three formations combined commonly exceed 1 km thickness. These are alluvial plain to alluvial fan deposits, and are consequently difficult to date, and contain numerous minor erosion surfaces and
periods of non-deposition. Hence episodes of rift quiescence that correspond with periods of inversion in the
southern part of the basin could easily be hidden within this type of poorly dated, continental stratigraphy.

Post-rift subsidence

The Malay, North Malay, and Pattani Basins are characterized by a Late Eocene(?) to Oligocene syn-

rift section and a Neogene post-rift section. The post rift section in the Pattani Basin exceeds 6 km thickness
in some places, and 8 km in the Malay Basin (Morley and Westaway, 2006; Fig. 14). Both the thickness and
subsidence rates are much greater than failed rifts with sag basins such as the North Sea and Anza Graben
(part of the Central African Rift System in Kenya), (Fig. 15). Note that the Anza Graben shows large, rapid
subsidence in the syn-rift phase, but minor subsidence in the post-rift phase. Also, for comparison subsidence curves from three classes of basin (data from Xie and Heller, 2009) are shown, plotted at the same scale
as the other basins. The only basins that compare in terms of subsidence rate are strike-slip basins, but it is
important to emphasize that the post-rift subsidence patterns of the Pattani and Malay basins are not fault
controlled, they are deep, saucer-shaped basins that have subsided rapidly over a broad area. In the Pattani

Basin minor normal faults are developed (but they are clearly not basin-bounding faults), while in the Malay Basin inversion anticlines grew during post-rift subsidence. The subsidence is also associated with high
heat flow (up to 110 mWm2) and geothermal gradients (up to 6C/100 m), (Morley and Westaway, 2006).
Consequently despite being about 23 my into the thermal subsidence phase, the basins are associated with
temperatures more typical of rift basins than post-rift basins. One way to try and explain the high subsidence rate is to begin the post-rift phase with a large accommodation space that needs filling (i.e. considerable
water depths), but that is impossible for these basins, the syn-rift fill is continental, while the post-rift fill is
predominantly fluvio-deltaic to marginal marine. Deposition has always been close to sea level, and even
today the greatest water depth in the Gulf of Thailand is only 80 m.

The basin depositional style is predominantly aggradational, and subsidence has kept pace with

sedimentation during deposition. Consequently the basins superficially look as if sediment loading has
driven basin subsidence. However, basic isostatic principals dictate that sediment loading cannot drive basin
subsidence to the extent required. As reviewed by Morley and Westaway (2006) the main explanations are
either extreme syn-rift crustal extension (Madon and Watts, 1998), or crustal flow away from the basinal
area during post-rift subsidence.

Huismans and Beaumont (2008) discussed a general model for passive margin development, where

some basins were undergoing sag basin formation, while adjacent basins underwent syn-rift, fault controlled
subsidence. Possibly a similar model in the Gulf of Thailand could be applied where crustal flow from
beneath the Pattani and North Malay basins was triggered by flow into the actively extending areas (Fig.
2). Gravity data over the Gulf of Thailand shows the Pattani and North Malay basins have a small negative
gravity anomaly, while the basement high areas have a larger negative gravity anomaly (Milsom, 2011), i.e.
uplift of the mantle (thinning of the crust) beneath the rifts has been so large as to dominate the expected
large negative gravity anomaly associated with a thick sedimentary basin. The opposite situation characterizes the western gulf, where large negative anomalies are associated with the sedimentary basins, and indicate crustal thinning is moderate (Milsom, 2011). Crustal flow to the west could help explain the pattern of
gravity anomalies in the gulf.

One effect of the rapid post-rift subsidence is the creation of high overpressures in the thicker parts

of the basin that can be a drilling hazard. The overpressure, which can attain a pore fluid pressure ratio of up
to 0.9, appears to be caused not only by disequilibrium compaction, but also by gas generation within coalrich sequences within the Middle Miocene (Tingay et al., 2013).
Post-rift extensional faults

On seismic reflection data, the syn-rift section of the Pattani Basin shows a clear half-graben style,

with predominantly N-S trending faults (Watcharanantakul and Morley, 2000; Morley and Westaway, 2006).
The North Malay basin shows a mixture of NW-SE and N-S trending faults syn-rift faults that bound half
graben (Morley et al., 2011). In the deepest parts of the basins the post-rift section is sufficiently thick that
the syn-rift is not imaged on 6 seconds two way travel time seismic data. The major syn-rift faults can have
throws of several kilometres. There are no faults of such magnitude that affect the post-rift section; however, anastomosing zones of low-displacement normal faults are very well developed. A seismic reflection
package characterized by continuous, high amplitude reflections forms a distinctive zone that separates the
syn-rift and post-rift sections. Known as the K-shales, this zone corresponds with Late Oligocene-Early
Miocene organic-rich lacustrine shales. Most of the faults visible on seismic data have throws in the order
of 10s metres, with the very infrequent, largest faults showing displacements up to about 400 metres. The
faults tend to form convergent dipping conjugate sets. These small faults set up numerous, but small traps,
where predominantly channelized sands trending at a high angle to the fault zone are laterally sealed by the
fault (Rigo de Rhigi et al., 2003; Racey, 2011). The sands are predominantly filled with gas or gas and condensate. The sands are stacked against the fault, and gas filled sands can alternate with sands with low gas
saturations, or water filled sands. Alternations of gas and water bearing sands can occur seven or eight times
in a well. The timing of fault activity is highly variable, with some faults terminating close to the sea floor,
while others ceased activity during the Miocene. Reactivation of the faults is a significant factor in breaching
traps and leaking fluids. Evidence for this fluid movement is seen in the mixing of gases of different thermal maturity in traps along the fault, and in the wide distributions of gas saturations in adjacent reservoirs.
Rarely are reservoirs filled to spill point.

Faults imaged on 3D seismic data in the post-rift basins of Thailand provide excellent examples

of multiple fault linkage geometries (Kornsawan & Morley 2002; Rigo de Rhigi et al. 2003; Morley et al.
2004). Contour patterns of the faults frequently show displacement maxima in the post-rift section, with displacement dying out downwards as well as upwards, indicating that the faults nucleated within the post-rift
section (Kornsawan & Morley 2002; Morley et al. 2007a) (Fig. 16). Some syn-rift faults have propagated
upwards into the post-rift section and there are examples of vertical linkage between syn-rift and post-rift
faults (Fig. 16a, b). Very long, low displacement faults are present in the post rift section (Fig. 16d,e). In the
Pattani Basin these faults tend to strike N-S, whilst in the North Malay Basin they strike NW-SE (Morley
et al. 2004) (Fig. 17b,e). These long-low displacement faults tend to link and align along underlying synrift faults and give rise to a long fault with multiple along-strike displacement highs and lows indicative of
previously isolated faults that have linked (Figs. 16b1, 17d,e). Some of these faults are 40-80 km long and

may be composed of 20-30 linked faults that were initially 1-4 km long, and may have maximum throws of
only 100-300 m. This mixture of low-displacement short and long faults produces a considerable spread of
points on a length-displacement graph (Fig. 18). The fault pattern is particularly well developed in the North
Malay Basin where short N-S striking fault segments curve to join long NW-SE trending faults (Leo 1997;
Morley et al. 2004). The long faults display significant under-displacement for their length (length: displacement ratios up to ~ 300:1), when compared with typical fault sets worldwide, (length: displacement ratios
around 10-20:1, e.g. Walsh & Watterson 1988; Dawers et al. 1993) (Fig. 18). For comparison the syn-rift
fault displacement-length data from the Mae Moh Mine (Northern Thailand) is plotted with the North Malay
Basin data (Fig. 18). The syn-rift faults display scatter in displacement: length ratio due to fault linkage, but
clearly the post-rift faults show much lower length: displacement ratios than the syn-rift faults.
The faults of the Pattani and North Malay basins represent an unusual style and occurrence of fault.
Unusual because they are systematic, low-displacement faults developed in a post-rift basin. The style is
atypical because the faults are conjugate in cross-section, but whereas ideal conjugate faults have sub-parallel trends in map view, the post-rift fault sets are commonly composed of two or three important directions
due to underlying syn-rift fabrics or local rotation of stresses (Morley et al., 2004; Tingay et al., 2010). Directions related to pre-existing (syn-rift) faults tend to give rise to very long, low-displacement faults. These
faults have trapped gas in very numerous, but small accumulations. Initially the accumulations were considered non-commercial, but Unocal were able to exploit them by the early application of 3D seismic technology to place deviated wells on stacked reservoir targets within 70-170 m of the fault, and by rapidly drilling
about 150 development wells per year, at very low cost (~0.8 million USD/well; Valusek, 1998).
The author is unaware of other examples of post-rift basins with the conjugate fault style displayed
by the Pattani and North Malay Basins. The faults have developed during the Middle Miocene to Present,
but the causes of their formation and activity remains uncertain. Tingay et al. (2010) suggested that the faults
formed episodically in response to strain imposed on the Gulf of Thailand from large earthquakes on the
Sumatra-Andaman subduction zone. This interpretation follows from observations of the considerable strain
gradient across the gulf arising from co- and post-seismic motions associated with the 2004 Sumatra earthquake. Whether this is a feasible mechanism remains to be determined.

Tectonics underlying the anomalous characteristics

The rift basins of Thailand are failed rifts within continental crust, but they lie within a distinctly

different tectonic setting from the failed rifts used for comparison in this paper. The Gulf of Suez, North
Sea, Central African and East African rifts all lie within old stable continental crust and all, except the Gulf

of Suez (failed propagating tip), are at least in part aulacogens branching off passive margins. In contrast,
Sundaland is surrounded by subduction zones, and has also undergone the collision of India with Eurasia,
and Australia with Indonesia. The Sundaland crust began growing by accretion of various terranes during the
Indosinian orogeny in the Triassic, and the process is continuing today (Hall, 2012). Hence, the crust is full
of relatively young heterogeneities at a variety of scales. In many places it is heavily intruded by TriassicEocene granitic plutons that contribute radiogenic heat both to the crystalline crust, and radiogenic sediments to basins (Morley and Westaway, 2006). Subduction processes also enhance the mantle heat flow into
the crust (Hyndman et al., Hall and Morley, 2004). Hence, in many places Sundaland crust is hot and weak
and consequently tectonics are much more dynamic in than in the other regions of failed rifts (Hall, 2011).
Conversely, there is no indication of mantle plume activity affecting Thailand.

The tectonic setting feeds back into the five anomalous features discussed above in a number of

ways. The highly variable location and timing of inversion, and the varied onset and termination of rifting
reflects a rapidly evolving stress field both temporally and spatially. The modern stress field, unlike most
continental areas is not dominantly aligned with the maximum horizontal stress parallel to absolute plate
motion (Richardson, 1992). Instead it is distinctly mis-aligned (Fig. 19; Tingay et al. 2010b ). This misalignment is a consequence of forces arising from multiple plate boundaries, plus intraplate sources of stress,
which have produced a more complex stress pattern than is typical of continental areas. Variations in stress
magnitude and orientation as the plate boundaries have evolved can help explain why alternation between
extension and inversion is so varied (e.g. Hall and Morley, 2004; Pubellier and Morley, in press).

The low-angle normal faults may reflect both the hot, weak, nature of the crust, and the effects of

pre-existing fabrics related to the collision history. The rapid post-rift subsidence can be related to two key
factors, a rapidly eroding sediment source area (largely local to Thailand; Hall and Morley, 2004), which in
particular reflects the tectonically active western region (Shan Plateau area), and also hot crustal conditions
that were conducive for rapid subsidence perhaps in response to lower crustal flow entirely related to postrift subsidence (Morley and Westaway, 2006), or where post-rift subsidence in the Pattani and Malay Basins
is related to crustal flow in response to non-uniform extension in adjacent active rift basins.
Conclusions

The intra-cratonic, supra-subduction zone setting of the Thailand rift basins have caused the rifts to

evolve in a number of ways that are different from other intra-cratonic rifts developed in older continental
lithosphere. The key differences focused on in this paper are the widespread occurrence of low-angle normal faults, mixed with higher-angle faults; basin inversion alternating with rifting; diachronous initiation

and cessation of rifting; rapid post-rift subsidence and the presence of extensive, low-displacement post-rift
fault. These characteristics are related to the presence of hot, weak continental lithosphere, a rapidly evolving stress regime due to the rapid evolution of the plate boundaries during the Cenozoic, and the history of
subduction and accretion (Fig. 20).
The impact of tectonics and the resulting structural and depositional characteristics is summarized in
Figure 20. Low-angle normal faults in general reduce the likelihood of source rocks developing, but in the
Sirikit Field resulted in fluvio-deltaic reservoirs interfingering with the northern margin of a lake controlled
by higher-angle faults. Basin inversion alternating with syn-rift section has been detrimental to the prospectivity of the southern half of the Phitsanulok Basin by destroying syn-rift structure, and not providing any
significant closed, inversion related structure, and halting hydrocarbon maturation early in the basin history.
The diachronous timing of basin development has been good for the eastern Gulf of Thailand basins where a
long history of rapid post-rift subsidence has resulted in deposition and maturation of source rocks, and kept
depositional facies belts quite stable, and aggradational in the post-rift section. Diachronous timing has not
been good for the younger basins that lack the post-rift section, and the greater variety of petroleum systems
and traps that post-rift basins offer. The origin of the post-rift faults is uncertain, but they are influenced by
pre-existing fabrics, and are in rapidly subsiding basins, which are probably a consequence of hot, weak
crust. They are the most significant hydrocarbon traps in the Pattani and North Malay basins, and provide
important migration pathways for fluids in the basin. They represent a structural style not usually found in
post-rift basins.
Acknowledgements
I would like to thank PTTEP for funding and data. Mark Tingay and Ian Watkinson are thanked for helpful
and constructive reviews.
Figures
Figure 1. Regional location map showing the main Cenozoic basins and tectonic features. Modified from
Searle and Morley (2011).
Figure 2. Regional cross-section through the southern Thailand rift basins, see Fig. 1 for location.
Figure 3 Regional map of Thailands Cenozoic basins and structures. Modified from Morley et al. (2011).

Figure 4 Generalized charts of Cenozoic structural activity in key rift basins in Thailand.
Figure 5. Schematic illustration of the modes of aulacogene development, A1 and A2), propagation inland
of the continental rift system (e.g North Sea); B) Simultaneous development of inland rift associated with
mantle plume, with a second mantle plume at the triple junction (e.g. Eastern Branch of East African Rift);
C) Simultaneous development of isolated rift basins along the length of the aulacogene, followed by small
local propagation to link up the basins (e.g. Central African Rift System).
Figure 6 Recoverable reserves for three categories of rift-related hydrocarbon province, A = simple rifts, B
= rifts with post-rift sag basins, and C = passive margins. Some provinces on passive margins were excluded
because they are much more related to large deltaic systems, independent of the syn-rift setting (these include the Nile and Niger Deltas and the Gulf of Mexico). West Siberia is excluded from post-rift sag basins
because it is such a large, and anomalous basin, and rifting may not be the only significant driver of basin
formation. Chart A = present known produced and unproduced reserves. Chart B = chart A plus predicted
P50 undiscovered resources, the changes largely reflect future deepwater exploration of the Atlantic passive
margins and how that may increase the relative contribution of passive margins to the total amount of hydrocarbons discovered in rift-related basins.
Figure 7 Example of a large, low-angle normal fault from the Mergui Basin. The fault plane is characterized
by high amplitude reflections down to a depth of about 12 km. The fault plane is low-angled and relatively
straight for most of its length. It does not appear to have been rotated significantly.
Figure 8 Models for the development of low-angle normal faults in response to different initial crustal conditions (temperature, thickness) according to Whitney et al. (2013).
Figure 9 Regional cross section from the Chiang Mai Basin across Doi Inthanon, illustrating the main structural and stratigraphic features, based on outcrop and seismic reflection data, and structural relationships
described by Dunning et al. (1995), Barr et al. (2002) and Morley (2009), apatite fission track ages from
Upton (1999). See Fig. 3 for location. b. Regional crustal-scale section across northern Thailand, the deep
crustal structure is mostly inferred from Bouguer gravity data. c. Sketch of the tectonic setting of central and
northern Thailand during the early Jurassic, at the end of the Indosinian Orogeny. The approximate location
of section a) is located to illustrate the origin of the key rock units (Sibumasu para- and ortho- gneisses), Pa-

laeozoic rocks above the Lower Detachment represent the Inthanon Zone. Redrawn from Sone & Metcalfe
(2008).
Figure 10 A) General structural map of the Phitsanulok Basin. B) Structural sketch map of inverted transfer
zone in southern part of the Phitsanulok Basin.
Figure 11 Cross-section through the Sirikit Field (see Fig. 10 for location). Based on Morley et al. (2007b).
Figure 12 Paleogeography of the Phitsanulok basin during the Late Oligocene-Middle Miocene, largely

based on unpublished maps made by Thai Shell in 1988. Maps based upon well data, supplemented by
2D seismic where sedimentary environment can be inferred from seismic character. Outlines of major
faults are shown.

Figure 13 A1), B1) Uninterpreted, and A2), B2) interpreted seismic lines showing inversion structures in the
southern Phitsanulok Basin. See Fig. 10B for location.
Figure 14 Cross-section through the Malay Basin, redrawn from Madon (2007).
Figure 15 Comparison of subsidence rates from different basin types (based in part on Xie and Heller, 2009,
and Morley and Westaway, 2006).
Figure 16 Schematic illustration of the way post-rift faults in the Pattani and North Malay basins interact.
a1, b1, c1 diagrams of displacement contours on fault surface, for fault i, darker grey indicates higher
displacement. a2, b2, c2 cross sections of post-rift conjugate fault systems, most of the faults have nucleated within post-rift section, but in cases a and b fault i has a relationship with syn-rift faults. a1. Fault
i is a syn-rift fault that has reactivated during the post-rift stage and propagated upwards into the post-rift
section, displacement dies out upwards and laterally. a2. cross-section showing large syn-rift offset on fault
i and much smaller post-rift offset. b1. Several low-displacement faults have nucleated within the post-rift
section but have been influenced (in location, dip and strike direction) by the deeper syn-rift fault (probably
due to local stress rotation) and linked with it. The single large displacement maxima on the syn-rift fault is
overlain by multiple, smaller magnitude displacement maxima in the post-rift faults. The map view displays
of fault patterns a,b1 and a,b2 are labelled to indicate that both geometries can occur for cases a and b.
a,b1. The syn-rift fault lies parallel to the regional extension direction because it is a large-displacement
syn-rift fault it is much longer than the lower-displacement post-rift faults. This is reflected in the length
of fault i. Note the conjugate faults tend to undergo systematic changes in dominant dip direction passing

along strike, this produces a series of transfer zones where the tips of the opposite-dipping faults slightly
overlap. In the Gulf of Thailand these features are informally called graben shifts. a,b2. the syn-rift fault
trends obliquely to the post-rift stress field, consequently fault i forms a long oblique fault trend relative to
the shorter post-rift faults. Post-rift faults with approximately similar dip direction to the syn-rift fault tend
to curve and join it. Post-rift faults dipping in the opposite direction cannot join the syn-rift trend, hence tend
to be short, isolated faults. c1. Fault i is one of a number of similar post-rift faults with no link to any synrift structure, displacement maxima is in the post-rift section and the fault dies out in all directions from the
centre. c2. A basement high nucleates the occurrence of the post-rift conjugate faults, no syn-rift faults are
present. The faults die out downwards. c3. Conjugate normal fault sets, with no dominant fault trend, reflecting the absence of long syn-rift faults. From Morley et al. (2011).
Figure 17 Illustration of syn-rift and post-rift fault patterns in the eastern Gulf of Thailand. a) Regional synrift fault map of the Pattani and Khmer basins based on 2D and 3D seismic reflection data, illustrating fault
patterns in the Late Oligocene-Early Miocene syn-rift section, and b) Detail of post-rift fault swarm patterns
from the Northern Pattani basin (redrawn from Rigo de Rhigi et al. 2002). c) The post-rift faults tend to form
curvi-linear trends of convergent conjugate faults with 10s meters to a few hundred metres displacement.
These trends often appear to be guided by underlying syn-rift faults which localize long, low-displacement
post rift faults, as illustrated for the Tantawan graben. d) Time structure map fo the Tantawan graben (redrawn
from Rigo de Rhigi et al. 2002), absolute scale not shown (darker colours = deeper time-depths). Many of the
post-rift faults tend to have different strikes from the underlying syn-rift faults but join, or splay off trends
controlled by the syn-rift faults. e) Example of fault map view geometry in the post-rift section from the North
Malay basin (Morley et al. 2004). In d) and e) the black faults dip E to NE, while the light grey faults dip to
the W to SW. (From Tingay et al., 2010a).
Figure 18 Length-displacement graph for post-rift faults in the North Malay Basin (faults mapped from 3D
seismic reflection data), compared with plots for syn-rift faults in Mae Moh mine (Morley and Wonganan
2000).
Figure 19 Distribution of misfit between maximum horizontal stress orientations and absolute plate motion
in the Sunda Plate (A-C quality in light grey, D quality in dark grey. From Tingay et al. (2010a).
Figure 20. Relationships between tectonic setting, structures and the hydrocarbon system of the Thailand

rift basins.

References
Abdulbaset, M. A., van Wees, J.-D., van Dijk, P. M., and Cloetingh, S. A. P. L., 2008. Tectonics and subsidence evolution of the Sirt Basin, Libya. AAPG Bulletin, 92, 993-1027.
Barr, S. M., Macdonald, A. S., Miller, B. V., Reynolds, P. H., Rhodes, B. P. & Yokart, B. 2002. New
U-Pb and Ar/Ar ages from the Doi Inthanon and Doi Suthep metamorphic core complexes, Northwestern Thailand. In: Symposium on Geology of Thailand, 26-31 August 2002, Department of Mineral Resources, Bangkok, Thailand, 284-308.
Bosworth, W., Huchon, P., and McClay, K., 2005. Red Sea and Gulf of Aden Basins. Journal of African
Earth Sciences, 43, 334-378.
Buck, W. R., 1991. Modes of continental lithospheric extension. Journal of Geophysical Research, 96,
20,161-20,178.
Burke, K., 1977. Aulacogens and Continental Breakup. Annual Review of Earth and Planetary Sciences, 5,
371-396.
Dunning, G. R., A. S. Macdonald, and S. M. Barr, 1995, Zircon and monazite U-Pb dating of the Doi
Inthanon core complex, northern Thailand: Implications for extension within the Indosinian Orogen.
Tectonophysics, 251, 197 213, doi:10.1016/0040-1951(95)00037-2.
Ebinger, C., 1989. Tectonic development of the Western Branch of the East African Rift System. Geological
Society of America Bulletin, 101, 885-903.
Ebinger, C. J., Bechtel, T. D., Forsyth, D. W. and Bowin, C. O., 1989. Effective elastic thickness beneath
the East African and Afar Plateaux and dynamic compensation of the uplifts. Journal of Geophysical
Research, 94, 2883-2901.
Fairhead, J. D., Green, C. M., Masterton, S. M. and Guiraud, R., 2013. The role that plate tectonics, inferred stress changes and stratigraphic unconformities have on the evolution of the West and Central
African Rift System and the Atlantic continental margins. Tectonophysics, 594, 118-127.
Flint, S., Stewart, D. J., Hyde, T., Gevers, C. A., Dubrule, O. R. F. & Van Riessen, E. D. 1988. Aspects of
reservoir geology and production behaviour of Sirikit Oil Field, Thailand: an integrated study using
well and 3-D seismic data. American Association of Petroleum Geologists Bulletin, 72, 1254-1268.
Fossen, H., Odinsen, T., Faerseth, R. B., and Gabrielsen, R. H., 2000. Detachments and low-angle faults
in the northern North Sea rift system. In: Nottvedt, A., (ed.), Dynamics of the Norwegian Margin,
Geological Society of London, Special Publications, 167, 105-131.

Foster, D., Gleadow, A., 1992. Morphotectonic evolution of rift-margin mountains in central Kenya. Earth
Planetary Science Letters, 113, 157-171.
Foster, D. and Gleadow, A., 1996. Sructural framework and denudation history of the flanks of the Kenya
rift. Tectonics, 15, 258-271.
Genik, G.J., 1993. Petroleum Geology of cretaceoustertiary rift basins in Niger, Chad, and Central African
Republic. American Association of Petroleum Geologists Bulletin, 77, 14051434.
Kornsawan, A. and Morley, C. K., 2002. The origin and evolution of complex transfer zones (graben shits)
in conjugate fault systems around the Funan Field, Pattani Basin, Gulf of Thailand. Journal of Structural Geology, 24, 435-499.
Kyrkjebo, R., Gabrielsen, R. H. and Faleide, J. L., 2004. Jurassic-Cretaceous synrift-post-rift transition of
the northern North Sea. Journal of the Geological Society, 161, 1-17.
Lyngsie, S. B., 2007. Continental sutures and their influence on rifting in the North Sea. Ph.D. Thesis, Geophysical Laboratory, Geological Institute, University of Denmark, 64 pp.
Hall, R. & Morley, C.K. 2004. Sundaland Basins, In: Clift, P. Wang, P., Kuhnt, W. & Hayes, D. (eds).
Continental-ocean interactions within East Asian Marginal Seas, AGU Special Publication, 149, 5587.
Hall, R., 2011. Australia-SE Asia collision: plate tectonics and crustal flow. In: Hall, R., Cottam, M.A. &
Wilson, M. E. J. (eds.) The SE Asian Gateway: History and Tectonics of the Australia-Asia Collision.
Geological Society, London, Special Publications, 355, 75-109. DOI: 10.1144/SP355.5 0305-8719
Hall, R., 2012. Late Jurassic-Cenozoic reconstructions of the Indonesian region and the Indian Ocean. Tectonophysics, 570-571, 1-41.
Hendrie, D. B., Kusznir, N. J., Morley, C. K., and Ebinger, C. J., 1994. Cenozoic extension in northern
Kenya: a quantitative model of rift basin development in the Turkana region. Tectonophysics, 236,
409-438.
Huismans, R. S. and Beaumont, C., 2008. Complex rifted continental margins explained by dynamical models of depth-dependent lithospheric extension. Geology, 36, 163-166.
Hyndman, R. D., Currie, C. A., Mazzotti, S. P., 2005. Subduction zone backarcs, mobile belts, and orogenic
heat. GSA Today, 15, 7 pp., doi: 10:1130/1052-5173(2005)015<4:SZBMBA>2.0.C);2
Jackson, J.A., White, N. J., Garfunkel, Z., and Anderson, H., 1988. Relations between normal-fault geometry, tilting and vertical motions in extensional terranes: an example from the southern Gulf of Suez.

Journal of Structural Geology, 10, 155-170.


Keller, R., Wendlandt, R. F. and Bott, M. H. P., 2006. Chapter 13 West and central African rift system.
Developments in Geotectonics, 25, 437-449.
Koehn, D., Aanyu, K., Haines, S., and Sachau, T., 2008. Rift nucleation, rift propagation and the creation of
basement micro-plates within active rifts. Tectonophysics, 458, 105-116.
Kinabo, B. D., Hogan, J. P., Atekwana, E. A., Abdelsalam, M. G., and Modisi, M. P., 2008. Fault growth
and propagation during incipient continental rifting: Insights from a combined aeromagnetic and
Shuttle Radar Topography Mission digital elevation model investigation of the Okavango Rift Zone,
northwest Botswana. Tectonics, 27, doi: 10.1029/2007TC002154.
Kusznir, N. J., Roberts, A. M., and Morley, C. K., 1995. Forward and reverse modeling of rift basin formation. Geological Society Special Publication, 80, 33-56.
Lacassin R., Hinthong, C.K., Siribhakdi, S. Chauviroj, A., Charoenravat, H., Maluski, P. H., Leloup, P.H.
& Tapponnier P. 1997. Tertiary diachronic extrusion and deformation of western Indochina: structure
and 40Ar/39Ar evidence from NW Thailand. Journal of Geophysical Research, 102 B5, 10013-10037.
Le Gall, B., Gernigon, L., Rolet, J., Ebinger, C., Gloaguen, R., Nilsen, O., Dypvik, H., Deffontaines, B.
and

Mruma, A., 2004. Neogene-Holocene rift propagation in Central Tanzania: Morphostructural

and aeromagnetic evidence from the Kilombero area. Geological Society of America Bulletin, 116,
490-510.
Logan, P., Curd, S., Downie, B., Weston, J., and Shaw, D., 2009. Exploration on the Frontier: Towards an
Understanding of the Albert Basin. Search and Discovery Article 10192, 9pp.
Macdonald, A. S., Barr, S. M., Dunning, G. R. & Yaowanoiyothin, W. 1993. The Doi Inthanon metamorphic core complex in NW Thailand: age and significance. Journal of Southeast Asian Earth Science,
8, 117-125.
Macdonald, A.S., Barr, S.M., Miller, B.V., Reynolds, P.H., Rhodes, B.P., Yokart, B., 2010. PTt constraints on the development of the Doi Inthanon metamorphic core complex domain and implications
for the evolution of the western gneiss belt, northern Thailand. Journal of Asian Earth Science, 37,
82104, doi:10.1016/j.jseaes.2009.07.010.
Macgregor, D. S., 1995. Hydrocarbon habitat and classification of inverted rift basins. In: Buchanan, J. G.
and Buchanan, P. G., (eds.), Basin Inversion, Geological Society of London, Special Publications,
88, 83-93.
Madon, M., 2007. Overpressure development in rift basins: an example from the Malay Basin, offshore Pen-

insular Malaysia. Petroleum Geoscience, 13, 169-180.


Madon, M.B. and Watts, A. B., 1998. Gravity anomalies, subsidence history and the tectonic evolution of
the Malay and Penyu basins (offshore Peninsula Malaysia). Basin Research, 10, 375-392.
Magnavita, L. P., Da Silva, R. R., Sanches, C. P., 2005. Guia de campo da bacia do Reconcavo, NE do Brasil (Field trip guide of the Reconcavo basin, NE Brazil). Boletim de Geociencias da Petrobras, 13,
301-333.
Milsom, J., 2011. Chapter 19, Regional Geophysics. In: Ridd, M. F.,Barber, A. J. & Crow, M. J. The Geology of Thailand. Geological Society, London, 493506
Morley, C. K., 1989. Extension, detachments and sedimentation in continental rifts (with particular reference to East Africa), Tectonics, 8, 1175-1192.
Morley, C. K., 1999, Comparison of hydrocarbon prospectivity in rift systems. in: Morley, C. K. (ed.) Geoscience of Rift Systems-Evolution of East Africa. AAPG Studies in Geology, 44, 233-242.
Morley, C. K. 2001. Combined escape tectonics and subduction rollback-back arc extension: a model for
the evolution of Cenozoic rift basins in Thailand, Malaysia and Laos. Journal of the Geological Society of London, 158, 461-474.

Morley, C. K. 2002a. Tectonic settings of continental extensional provinces and their impact on sedimentation and hydrocarbon prospectivity. In: Renaut, R. W. & Ashley, G. M. (eds.). Sedimentation in
Continental Rifts. Society for Sedimentary Geology Special Publication, 73, 25-56.
Morley, C. K. 2002b. A tectonic model for the Tertiary evolution of strike-slip faults and rift basins in SE
Asia. Tectonophysics, 347, 189-215.
Morley, C. K. 2007. Variations in Late Cenozoic-Recent strike-slip and oblique-extensional geometries,
within Indochina: the influence of pre-existing fabrics. Journal of Structural Geology, 29, 36-58.
Morley, C. K. 2009. Geometry and evolution of low-angle normal faults (LANF) within a Cenozoic highangle rift system, Thailand: Implications for sedimentology and the mechanisms of LANF development. Tectonics, 28, doi: 10.1029/2007TC002202
Morley, C. K. & Racey, A. 2011. Tertiary stratigraphy. In: Ridd, M. F.,Barber, A. J. & Crow, M. J. The
Geology of Thailand. Geological Society, London, 223271
.Morley, C. K. & Westaway, R. 2006. Subsidence in the super-deep Pattani and Malay basins of Southeast
Asia: a coupled model incorporating lower-crustal flow in response to post-rift sediment loading.

Basin Research, 18, 51-84.


Morley, C. K. & Wonganan, N. 2000. Normal fault displacement characteristics, with particular reference
to synthetic transfer zones, Mae Moh Mine, Northern Thailand. Basin Research, 12, 1-22.
Morley, C. K, Nelson, R. A., Patton, T.L. and Munn, S. G., 1990. Transfer zones in the East Africa Rift
system and their relevance to hydrocarbon exploration in rifts. AAPG bulletin, 74, 1234-1253.
Morley, C. K., Sangkumarn, N., Hoon, T. B., Chonglakmani, C. & Lambiase, J. 2000. Structural evolution
of the Li Basin northern Thailand. Journal of the Geological Society of London, 157, 483-492.
Morley, C. K., Wonganan, N., Sankumarn, N., Hoon, T. B. Alief, A. & Simmons, M. 2001. Late Oligocene-Recent stress evolution in rift basins of Northern and Central Thailand: Implications for escape
tectonics. Tectonophysics, 334, 115-150.
Morley, C. K., Wonganan, N., Kornasawan, A., Phoosongsee, W., Haranya, C. & Pongwapee, S. 2004.
Activation of rift oblique and rift parallel pre-existing fabrics during extension and their effect on deformation style: Examples from the rifts of Thailand. Journal of Structural Geology, 26, 1803-1829.
Morley, C. K., Gabdi, S. & Seusutthiya, K. 2007a. Fault superimposition and linkage resulting from stress
changes during rifting: Examples from 3D seismic data, Phitsanulok Basin, Thailand. Journal of
Structural Geology, 29, 646-663.
Morley, C. K., Ionnikoff, Y., Pinyochon, N. & Seussutthiya, K. 2007b. Degradation of a footwall fault
block with hanging-wall fault propagation in a continental-lacustrine setting: How a new structural
model impacted field development plans, The Sirikit Field, Thailand. American Association of Petroleum Geologists Bulletin, 91, 1637-1661.
Morley, C. K., Charusiri, P. & Watkinson, I. 2011. Structural geology of Thailand during the Cenozoic. In:
Ridd, M. F., Barber, A. J. & Crow, M. J. (eds.), Geology of Thailand, Geological Society of London,
273-334.
Moucha, R., and Forte, A., M., 2011. Changes in African topography driven by mantle convection. Nature
Geoscience, 4, 707-712, doi: 10.1038/ngeo1235.
Nyblade, A. A., 2002. Crust and upper mantle structure in East Africa: Implications for the origin of Cenozoic rifting and volcanism and the formation of the magmatic rifted margin. Special Paper of the
Geological Society of America, 362, 15-26.
Pik, R., 2011. Geodynamics: East African on the rise. Nature Geoscience, 4, 660-661. Doi:10.1038/
nego1274.
Polachan, S., Pradidtan, S., Tongtaow, C., Janmaha, S., Intarawijtr, K. and Sangsuwan, C., 1991. Development of Cenozoic basins in Thailand. Marine and Petroleum Geology, 8, 84-97.

Racey, A. 2011. Chapter 13, Petroleum Geology. In: Ridd, M. F., Barber, A. J. & Crow, M. J. (eds.), Geology of Thailand, Geological Society of London, 352-392.
Rattey, R. P. and Hayward, A. B., 1993. Sequence stratigraphy of a failed rift system: The middle Jurassic
to early Cretaceous basin evolution of the Central and Northern North Sea. In: Parker, J. R. (ed.),
Petroleum Geology of Northwest Europe, Proceedings of the Fourth Conference, Geological Society,
London (1993), 215-249
Rigo de Rhigi, L., Baranowski, J., Chaikiturajai, C., Nelson, G,Wechsler, D. &Mattingly, G., 2003. Block
B8/32, Gulf of Thailand petroleum system and implementation of technology in field development.
Seapex Press, 6, 46-55.
Rhodes, B. P., Blum, J., and Devine, T., 2000, Structural development of the mid-Cenozoic Doi Suthep
metamorphic core complex and western Chaing Mai basin, northern Thailand. Journal of Asian
Earth Sciences, 18, 97 108, doi:10.1016/S1367-9120(99)00019-X.
Rhodes, B. P., Conejo, R., Benchawan, T. Titus, S., and Lawson, R., 2005, Palaeocurrents and provenance
of the Mae Rim Formation, northern Thailand: Implications for tectonic evolution of the Chiang Mai
Basin, Journal of the Geological Society, London, 162, 51 63, doi:10.1144/0016-764903-128.
Richardson, R. M., 1992. Ridge forces, absolute plate motions and the intraplate stress field. Journal of
Geophysical Research, 97, 11739-11748.
Roberts, E. M., Stevens, N. J., OConnor, P. M., Dirks, P. H. G. M., Gottfried, M. D., Clyde, W. C. ,
Armstrong, R. A., Kemp, A. I. S., and Hemming. S., 2012. Initiation of the western branch of the East
African Rift coeval with the eastern branch. Nature Geoscience, DOI: 10.1038/NGEO1432.
Sachau, T. and Koehn, D., 2010. Faulting of the lithosphere during extension and related rift-flank uplift: a
numerical study. International Journal of Earth Sciences, 99, 1619-1632.
Sone, M. & Metcalfe, I. 2008. Parallel Tethyan sutures in mainland Southeast Asia: New insights for
Palaeo-Tethys closure and implications for the Indosinian Orogeny. Comptes Rendus Geoscience,
340, 166-179.
Spiegel, C., Khon, B., Belton, D., Gleadow, A., 2007. Morphotectonic evolution of the central Kenya rift
flanks. Geology, 35, 427-430.
Tapponnier, P., Peltzer, G. & Armijo, R. 1986. On the mechanism of collison between India and Asia. In:
Coward, M.P. & Ries, A.C., (eds.). Collision Tectonics. Geological Society, London, Special Publications, 19, 115-157.

Tingay, M. R. P., Morley, C. K., Hillis, R. R. and Meyer, J., 2010a. Present-day stress orientation in Thailands basins. Journal of Structural Geology, 32, 235-248.
Tingay, M., Morley, C. K., King, R., Hillis, R., Coblenz, D., and Hall, R., 2010b. Present-day stress field
of Southeast Asia. Tectonophysics, 482, 92-104.
Upton, D. R. 1999. A regional fission track study of Thailand: implications for thermal history and denudation. Ph.D. thesis, University of London, 158 pp.

Valusek, J.E., 1999. Unocal Thailand slashes well planning cycle time. Oil and Gas Journal. 97, 45-48.
Wernicke, B., Axen, G. J., 1988, On the role of isostacy in the evolution of normal fault systems. Geology,
16, 848-851.
Watcharanantakul, R. & Morley, C. K. 2000. Syn-rift and post-rift modelling of the Pattani Basin, Thailand, evidence for a ramp-flat detachment. Marine and Petroleum Geology, 17, 937-958.
Whitney, D. L., Teyssier, C., Rey, P. and Buck, R. W., 2013. Continental and oceanic core complexes. 2013.
GSA Bulletin, 125, 273-298.

iao
d
Re

Sagaing Fault

ed
-R
Ri

ve

rS

he

ar

Qui
Nho

ra
m
Lin
e

MALAYSIAN
PENINSULA

W.
Natuna
Basin

at

Lup

ar L

ra

ma

ul

da

Fa

An

Spratly RiseDangerous
Grounds

Ba

Malay basin

Su

M = Mogok MCC, 26-21 Ma


DI-DS = Doi Inthanon-Doi Suthep MCC

sin

Ba

ug

ul

Fa

ong

L
uu

N.
Sumatra
Basin

Metamorphic core complex (MCC)


(Late Oligocene-Early Miocene)

n-S

Sumatra

Borneo

ine

Tarakan
Basin

Kutai Basin

Sunda
Shelf

ren
aT

atr

um

Approximate area of basin

Gulf of
Thailand

Fig. 2

HT = Himalayan Terrane
LCT = Lhasa-Coqen Terrane
TB = Tanggula Block
IYS = Indus Yarlung Suture

Bengal Basin (foredeep basin)

90

idge

n
Pi

East
AndamanMergui
Basin

South China
Seas oceanic
crust

ng R

ae

Andaman
Sea

Lanping-Siamo
Fold Belt and Khorat Plateau

iv
rl R
Pea
in
Bas
Hoang Sa
basin

Gulf
Dai Nui
Con Voi of Tonkin
Basin
Khorat
Plateau

Three Pagodas
Fault
15

outh

er M

DIDS

Indo-Burma Ranges
Central Basin

ne

Tr
o

Bengal Basin

Zo

an

INDIA

aw

Himalayas

an

Sh

HT

China Block

Tibetan Plateau rifts and


metamorphic core complexes

Al

IYS

TB

LCT

Pa
l

30

Approximate outline of Sundaland

ch

GPS derived velocity


20 mm/yr with respect to
Sundaland
(Simons et al. 2007)

500 km

Sundaland GPS station

105

Java

Fig. 1

East Andaman Basin


Sewell Rise

Mergui Basin

Peninsular
Thailand

Songkhla Basin

Pattani Basin

Depth (km)

10

Fig.
20
30

Fig. 2

Upper Crust

Lower Crust

100 km

Flow of crust?

Khmer Basin

Chiang Rai

PU

Pua

CRP

Chiang Rai-Payao

CM

Chiang Mai

MLA

Mae Lamao

Myanmar

Mae Lai

PH

WH

Phitsanulok

LP

Lampang

LI

MSR MC

Li

Pi

MC

CM

DI

Phrae

MS

Mae Ping
Fault Zone

Mae Sot

MSR

Mae Sariang

TH

Thoen Basin

PET

Phetchabun

Sakhon Nakhon

SKN

Nong Bua

NB

Lad Yao

LY

Nakhon Sawan

NS

Suphan Buri

SB

Ayutthaya

AY

Kamphaeng Saen

KS

Paknam

PN

Hua Hin

HH

N. Western

NW

Western

WE

Kra

NB

LY

NS

WH

SB
KS

South Pattani

CB

TB

HH

Khmer

Prachuap

Chumphon
Nakhon

NMB
NYP
EA

North Malay Basin

Surat Thani

Nong y Plong
East Andman
Cenozoic strike-slip
or normal fault

DI = Doi Inthanon
Cenozoic basin
Deformed granitegneiss that may
display Cenozoic
metamorphism and
migmatites
Cenozoic folds in
Khorat Group

Fig. 3

ge

Krabi

gui
Rid

Mergui

MA

MER

Cambodia
Possible early Cenozoic
migmatites in gneisses
exhumed along the
Three Pagodas Fault Zone

PN
NW

Khao Taphao
Khwam (low-angle
normal fault

GOT

12oN-

WE
CHP

NPA

Ko Samui
High

EA

Mer

Songkhla

SO
KB

AS

Mergui

MER

NK

aul

CHP

gF

PR

PT

non

KH

14oN-

Bangkok

PR

Ranong Fault
mylonites,
Khao PLai
Khlong Hin
Phao

16oN-

Thailand

EK

KB

Ra

SPA

KU

AY

SK

NYP

KP

PET

MLA
MS

Klo
Fau ng Ma
rui
lt

Central Pattani

CPA

WH = Western Highlands
NCB = North Central Region
CB = Central Basin
KP = Khorat Plateau
GOT = Gulf of Thailand
PT = Peninsular Thailand
AS = Andaman Sea

SKN

North Pattani

NPA

200 km

18oN-

East Kra

EK

Uttaradit
Fault Zone

LP MM PHR

PH

Hua-HinPran Buri
Mylonites,
northern
Ranong
Fault
Zone

Sakhon

SK

Laos

PU

TH

Outline of
Chao Phraya
Basin

Thon Buri

TB

CRP

100

LI

Three Pagodas
Fault zone

Khorat Udon

KU

Mae Chan
Fault

NCB

Mae Moh

PHR

l
102oE

CR

ML

Mae Chaem

MM

l
100oE

Chiang Mai Basin

Nan

ML

ST

l
98oE

Fang

KH

10oN-

CPA

Khanom
ST

NK
Khao
Phanom
mylonites

South

SO
SPA

Narathiwat High

CR

8oN-

NMB

6oN-

Malaysia

Western GOT

Eastern GOT

Miocene

NM

Kh

N-S

Ch

We

SB

North Central
Ph

Chi/L

PlioceneRecent

Late

Pa

Central

10

Middle

Oligocene

Early 20
Late
Early 30

Eocene

Late

40
Middle

Ma
Unconformity

Inversion

Period of
rapid post-rift
subsidence

Post-rift basin

Transitional syn-rift
to post-rift

Syn-rift

Fig. 4

A2

Aulacogen

Mantle
plume

A1
Mantle
plume

Triple junction

ive
Pass

gin
mar

Spreading centre

Fig. 5

Rift propagation
direction

MM

CM

Total 291 BBOE

Total 416 BBOE

C
B

A = simple rifts, B = rifts with post-rift sag basins, (excluding West Siberia)
C = passive margins (excluding large deltas)

Fig. 6

W
A

1.0

TWTT (s.)

2.0
3.0
4.0
5.0
6.0

B
1.0

Middle MioceneRecent

2.0
Early MioceneOligocene (syn-rift)

3.0
4.0

Pre-Rift

5.0
6.0

Fig. 7

22 average dip

5 km

Approximate depth
12 km

A. WARM CRUST

B. W HOT CRUST

Block-rotation model for core complex formation

Rolling-hinge model for core complex formation


1 2

Extensional basins

Moho

Incipient normal faulting, block rotation,


lower crustal flow; fault 1 stop slipping, fault 2 takes over
resulting in passive rotation of fault 1

Incipient normal faulting, block rotation,


and lower crustal flow
domino rotation
of upper crust blocks

Order of faulting

Flow of lower crust


Moho

Moho

Pre-extension Moho

4 5

Flow of lower crust

Pre-extension Moho

~ 50 km

C. HOTTEST CRUST
C1. Converging channel flow of low-viscosity
(partially molten) lower crust

C2. Core complex developed at edge of orogenic plateau


PLATEAU: THICK, HOT CRUST

FORELAND: COLD CRUST

Solidus
Moho

Moho
Pre-extension Moho

Fig. 8

Fixed boundary

Channel detachment

Flowing, partially molten crust


Pre-extension Moho

Upper crust
Lower crust
Transfer of thick crust
towards foreland

Doi Inthanon
Remnant Tertiary basins
(e.g. Mae Cham Mine)

Muscovite Ar/Ar
21 1 Ma

Depth (km)

Lower Detachment
Muscovite Ar/Ar
18 0.2 Ma
Apatite Fission Track
Chiang Mai Basin
CMLANF
central age 16 Ma 1Ma
Inversion anticline
and 14 Ma 1 Ma

?
5

? ?

? ?
5 km
Triassic granite

Doi Inthanon orthogneiss

Mesozoic
sedimentary rocks

Mae Klaeng Granite


(Oligocene)

Paragneiss

Palaeozoic metasedimentary rocks

-60
-70
-80
-90
-100
-110

shape of
gneiss dome?

Cenozoic syn-rift
sediments

Bouger gravity anomaly (mg.)

Mae Yuan Fault zone


Detachment fault
0

? ?

Mae Cham
Basin

Doi Inthanon

CMLANF

LD

Paragneiss

Orthogneiss

Mae Tha Fault

Chiang Mai Basin

Payao Fault

LD

10

20

30
km

Moho

v=h

25 km

Approximate location of Late Oligocene brittle-ductile transition/ Lower Detachent surface (LD)
Estimated present-day brittle-ductile transition

Predominantly Palaeozoic rocks, overlying the Lower Detachment (LD)

Early Jurassic

Palaeo-Tethys
Suture Zone

Meso-Tethys

Nan-Sa Kaeo
Back-arc suture
Sukhothai
Loei-Phetchabun
Arc
Volcanic Belt
Indochina

Sibumasu

Fig. 9

S-type
granitoids
Main-Range
Granitoid Province

I-type
Eastern Granitoid
Province

S-type
Western Granitoid
Province

Phichai Sub-basin

Uttaradit
Fault

Western
Boundary
Fault

Fig. 13A

Fig. 13B

Region of thin
syn-rift section
and widespread
inversion
Area of well
developed rift
basins

Sukhothai Sub-basin

Inverted
normal fault
G

Fig. 11

Sirikit Field

Dong Chat
Sub-Basin

B
Main normal
fault trends

Large normal fault

Lahan
Graben

5 km

Extension mostly
during Miocene,
Late MiocenePliocene inversion

Nong Bua
Sub-Basin

Mixed Miocene and


Late Oligocene extension
Late Miocene-Pliocene
inversion
Extension mostly OligoceneEarly Miocene, some
mid-late Miocene extension
alternating with inversion

Fig. 10

Inversion anticline
and syncline

10 km

1000

E
Sealing lacustrine shale
LKU-E17

Oil reservoir
Gas reservoir

LKU-R05
TMS

Depth (metres)

1500

Main Seal

K sands GWC

Spill point?

K sands OWC
K-Sands

Main Seal
TP

LKU-A05
?

L-sands

P sands OWC

LIS

LKU-E22
LKU-E03

M-sands
M sands OWC
Basal Seal

P-sands

2500

Fig. 11

UIS

L sands OWC

2000

200 m

LKU-A06

Late Oligocene

Early Miocene, Lower Chum Saen Fm,


Basal Seal times

Early Miocene, Lan Krabu,


M-sand times

?
?

A
Early Miocene, Lan Krabu,
M-sand times

High-angled normal falts

oundary

~45* dips

B
Western

20-30 fault zone dips

Fault

Western Boundary
Fault becomes
active

Small, high-angle
fault-bounded
depression

?
?
?

Early Miocene, Upper


Chum Saeng Fm,
Main Seal

Middle Miocene
Pratu Tao Fm.
Uttardit Fault zone
becomes active

10 km

Sirikit Field

E
Scree, alluvial fans, fan delta

Alluvial plain with


meandering rivers

Freshwater lake

Alluvial plain with


non-meandering rivers

Lacustrine delta-paralic sediments

Wells -control points for sedimentary


environment

FIg. 12

Two way travel time (s.)

A1
2

Two way travel time (s.)

Late -Middle\
Miocene

1
OligoceneMiddle Miocene

Pre-Cenozoic basement

1 km

Two way travel time (s.)

E
Pliocene-Holocene

A2

C1

C2

B1
0

Two way travel time (s.)

Minor second phase of inversion folding


unconformity

B2

Fig. 13

Two inversion events on this fault, first event by folding and erosion prior to yellow horizon
second even by uplift and erosion of brown unit, green horizon forms a local unconformity.

C3
E

km 0

ENE

WSW

10

Tok Bidan
Graben

Raya Graben

Pre-Cenozoic Basement

12
0

50
Oligocene

100

150

Lower Miocene

Middle Miocene

200 km
Late Miocene- Holocene

Fig. 14

225

200

150

Time (Ma)

100

50

0
1
2
3

7
8
9
10
11
12

Pattani basin
Malay basin

Strike-slip basins

Foreland basin

Passive margins

Depth (km)

Fig. 15

Post-rift Syn-rift

A1

B1

C1

Boundary between
syn-rift and postrift section

Post-rift

Post-rift

Syn-rift

Syn-rift
Pre-Rift

A2
A,B 1

Shmax

Fig. 16

i
Post-rift

Pre-Rift

Transfer zone
(graben shift)

B2
A,B 2

Pre-Rift

i
Shmax

C2
C3

Shmax

Transfer zone
(graben shift)

Detail of
Tantawan
Graben

b
Pattani
Basin

N
N Malay
Basin

Area of
map b
Offshore
Cambodia

Offshore
Thailand

Benchamas
Horst

c
Ko Kra Ridge

3500

North
Pattani Basin

Khmer Basin
3000

Ko Kra Ridge

Syn-rift faults
underlying
post-rift
fault zone

Joint
Claims
Area

Timestructure
map

Maliwan Graben

Central (main)
Pattani Basin
South

Jarmjuree

North Malay Basin


Southern
Pattani
Basin

2 km
Tantawan Graben

Narathaiwat High
Syn-rift normal faults
Rift basin depocentres

Basement highs

Ranong
Terrace

Peninsula Thailand

Surat High

10 km

50 km

2 km

Long, low-displacement
faults with multiple
fault linkage geometries

Fig. 17

35

Typical distribution,
e.g. stable North
America

Number of Indicators

30
25
20
15
10
5
0
-80

-60

-40

-20

20

40

60

80

Misfit with Maximum Horizontal Stress Azimuth (degrees)


Fig. 18

Displacement - D (m)

1000

100
-1

10

D=

10

2
D=

-1

-2

1
100

0
=1

Mae Moh Mine


(syn-rift)

-3

D=
1000

10

N. Malay Basin
(post-rift)
10000

Fault trace length- L (m)

Fig. 19

Fig. 11.44
Morley et al.

100000

Persistent subduction zone setting

Accretion

Thickening of crust-extensional collapse?


Pre-existing fabrics

Low angle normal


faults

Rapidly evolving stress regime

Basin inversion

Diachronous basin
timing

Hot, weak continental


lithosphere

Rapid post-rift
subsidence

Post rift faults

Trap
Source
Seal
Reservoir

Migration
Geo-hazard

Overpressure

Impact on hydrcarbon system

Fig. 20

Negative

Positive

Mixed

Das könnte Ihnen auch gefallen