Download as pdf or txt
Download as pdf or txt
You are on page 1of 51

Earth-Science Reviews 104 (2011) 41–91

Contents lists available at ScienceDirect

Earth-Science Reviews
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 / e a r s c i r ev

Deepwater fold and thrust belt classification, tectonics, structure and hydrocarbon
prospectivity: A review
C.K. Morley a,⁎, R. King b, R. Hillis c, M. Tingay b, G. Backe b
a
PTTEP, 27th Floor, ENCO Building, Soi 11, Vibhavadi-Rangsit Road, Chatuchak, Bangkok, Thailand, 10900
b
Centre for Tectonics, Resources and Exploration (TRaX), Australian School of Petroleum, University of Adelaide, SA 5005, Australia
c
Deep Exploration Technologies Cooperative Research Centre, c/o University of Adelaide, SA 5005, Australia

a r t i c l e i n f o a b s t r a c t

Article history: Deepwater fold and thrust belts (DWFTBs) are classified into near-field stress-driven Type 1 systems confined
Received 3 April 2010 to the sedimentary section, and Type 2 systems deformed by either far-field stresses alone, or mixed near- and
Accepted 28 September 2010 far-field stresses. DWFTBs can occur at all stages of the Wilson cycle up to early stage continent continent
Available online 4 November 2010
collision. Type 1 systems have either weak shale or salt detachments, they occur predominantly on passive
margins but can also be found in convergent-related areas such as the Mediterranean and N. Borneo.
Keywords:
deepwater fold and thrust belts
Examples include the Niger and Nile deltas, the west coast of Africa, and the Gulf of Mexico. Type 2 systems are
hydrocarbons subdivided on a tectonic setting basis into continent convergence zones and active margin DWFTBs. Continent
thrust convergence zones cover DWFTBs developed during continent–arc or continent–continent collision, and
fold those in a deepwater intracontinental setting (e.g. W. Sulawesi, Makassar Straits). Active margins include
accretionary prisms accretionary prisms and transform margins. The greatest variability in DWFTB structural style occurs between
deltas salt and shale detachments, and not between tectonic settings. Changes in fold amplitude and wavelength
salt tectonics appear to be more related to thickness of the sedimentary section than to DWFTB type. In comparison with
shale tectonics
shale, salt detachment DWFTBS display a lower critical wedge taper, more detachment folds, long and
growth faults
episodic duration of deformation and more variation in vergence. Structures unique to salt include canopies
continent collision
gravity-driven deformation and nappes. Accretionary prisms also standout from other DWFTBs due to their relatively long, continuous
near-field stress duration, rapid offshore propagation of the thrust front, and large amount of shortening. In terms of
far-field stress petroleum systems, many similar issues affect all DWFTBs, these include: the oceanward decrease in heat
Niger Delta flow, offshore increase in age of mature source rock, and causes of trap failure (e.g. leaky oblique and frontal
Borneo thrust faults, breach of top seal by fluid pipes). One major difference between Type 1 and Type 2 systems is
Gulf of Mexico reservoir rock. High quality, continent-derived, quartz-rich sandstones are generally prevalent in Type 1
Caspian Sea systems. More diagenetically reactive minerals derived from igneous and ophiolitic sources are commonly
Timor
present in Type 2 systems, or many are simply poor in well-developed turbidite sandstone units. However,
Trinidad
Barbados
some Type 2 systems, particularly those adjacent to active orogenic belts are partially sourced by high quality
continent-derived sandstones (e.g. NW Borneo, S. Caspian Sea, Columbus Basin). In some cases very high rates
of deposition in accretionary prisms adjacent to orogenic belts, coupled with uplift due to collision, results in
accretionary prism related fold belts that pass laterally from sub-aerial to deepwater conditions (e.g. S.
Caspian Sea, Indo-Burma Ranges). The six major hydrocarbon producing regions of DWFTBs worldwide (Gulf
of Mexico, Niger Delta, NW Borneo, Brazil, West Africa, S. Caspian Sea) stand out as differing from most other
DWFTBs in certain fundamental ways, particularly the very large volume of sediment deposited in the basins,
and/or the great thickness and extent of salt or overpressured shale sdetachments.
© 2010 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
2. Outline of the DWFTB classification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
2.1. Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
2.2. Stress-system terminology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

⁎ Corresponding author. Tel.: + 6625374264.


E-mail address: chrissmorley@gmail.com (C.K. Morley).

0012-8252/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.earscirev.2010.09.010
42 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

2.3. Are near-field stress systems confined to passive margins? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44


2.4. Tectonic setting: active margins and convergent zones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
2.5. Classification scheme. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3. The characteristics of salt and shale detachments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.1. Gravity sliding and gravity spreading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.2. Salt detachment zones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.3. Shale detachments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.4. Do thick mobile shale zones exist? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
4. Structural and petroleum system characteristics of the different types of DWFTB . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.1. Near-field stress-driven linked systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.1.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
4.1.2. Type 1a, DWFTBs associated with regional offshore-dipping detachments along overpressured shales . . . . . . . . . . . . . 54
4.1.3. Type 1a, DWFTBs associated with a widespread hinterland-dipping basal detachment zone (large deltas) . . . . . . . . . . . 56
4.1.4. Type 1b, DWFTBs associated with an oceanward-dipping salt detachment zone . . . . . . . . . . . . . . . . . . . . . . . . 58
4.2. Type 2, continental convergent tectonic setting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.2.1. Type 2a convergent zones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.2.2. Type 2bi weakly linked/unlinked DWFTBs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
4.3. Type 2bii active margins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.3.1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
4.3.2. Barbados . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
4.3.3. The Makran . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
4.3.4. Andaman subduction zone–Indo-Burma Ranges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.3.5. South Caspian Sea . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
5. Synthesis of DWFTB characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
5.1. Petroleum systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
5.1.1. Source rock types . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
5.1.2. Temperatures and hydrocarbon maturation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
5.1.3. Hydrocarbon migration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
5.1.4. Reservoir quality . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
5.1.5. Structural traps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81
5.1.6. Seal . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
5.2. Structural development in different tectonic settings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
6. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

1. Introduction theme of DWFTBs. The broad division of DWFTBs into passive and
active margins, shale vs. salt detachments, and gravity sliding vs.
There has been long-standing academic interest in deepwater fold and gravity spreading mechanisms has been proposed previously (e.g.
thrust belts (DWFTBs), particularly along subduction zones and early- Morley and Guerin, 1996; Rowan et al., 2004; Krueger and Gilbert,
stage collisional margins. Over the last two decades interest has surged 2009). This paper discusses refinement of this classification, because a
following advances in deepwater drilling technology by the oil industry. simple grouping into gravity-driven passive margins and lithospheric-
Deepwater exploration encompasses many different potential traps and stress driven, tectonically active margins (e.g. Rowan et al., 2004;
geological settings including DWFTBs, which have featured prominently Hamilton and De Vera, 2009) does not encompass all settings of
because they contain numerous large anticlines with associated hydro- DWFTBs. Instead, a classification system for DWFTBs is proposed
carbon seeps. At present, six areas are the main focus of deepwater based on the driving mechanism, detachment type and tectonic
development and production in DWFTBs, these are: the Gulf of Mexico, setting (Fig. 2).
Niger Delta, NW Borneo, the Brazilian Margin, West Africa (Angola, The justification for the classification scheme used in this paper is
Congo) and the South Caspian Sea (mostly shallow water, but extending addressed in Section 2. Then the nature of the detachment zones is
into deepwater) (Fig. 1). Exploration is being conducted in many more. discussed in depth, particularly for shale detachments, since their
The academic and industry activity mentioned above has produced variations exert a strong control on DEFTB structural style. Currently,
a wealth of information about DWFTBs that is not only important for the nature of mobile shale detachments is a subject of controversy,
high-grade areas in which very expensive exploration and develop- with some workers doubting whether thick mobile shale masses can
ment programs are being conducted, but also for fundamentally exist at depth in the subsurface. The fourth and largest part of the
modifying and improving our understanding of how passive margins, paper (Section 4) then describes the tectonic, structural geology and
and the early stages of orogenic belts develop. This paper is the first petroleum system characteristics of the different types of DWFTBs in
detailed review of the considerable body of published data on modern the context of the proposed classification scheme. The differences and
DWFTBs (not ancient, exhumed DWFTBs); it reviews the significant similarities in petroleum systems within the different types of DWFTB
characteristics of key deepwater fold and thrust belts, and their are discussed in Section 5.
petroleum systems. The DWFTBs are reviewed as a group because
despite the very different tectonic processes that can operate in the 2. Outline of the DWFTB classification
hinterlands of the DWFTBs, there are many similar characteristics to
the active DWFTB in all the different tectonic environments. The 2.1. Introduction
examples of DWFTBs in different settings are described in detail in
Section 4 because they either represent well-documented examples, Previous classification schemes have noted that DWFTBs can either
or they illustrate some of the significant variations found within the develop under the influence of gravity where deformation is limited to
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 43

Fig. 1. Digital elevation map (DEM) of the world, showing locations of deepwater fold and thrust belts, and their classification (see Fig. 2). Note: only some examples of accretionary
prisms are shown. 1 = McKenzie Delta, 2 = Cascadia accretionary prism, 3 = Perdio DWFTB, 4 = Mississippi Fan DWFTB, 5 = Mexican Ridges DWFTB, 3, 4 and 5 = Gulf of Mexico,
6 = Scotia Basin, 7 = Barbados accretionary prism, 8 = Columbus Basin, 9 = Amazon Delta, 10 = Para-Maranhao, 11 = Sergipe-Alagoas, 12 = Espirito Santo Basin, 13 = Campos Basin,
14 = Santo Basin, 15 = Pelotas Basin, 16 = Straits Gibraltar (Prerif'nappe'), 17 = Essaouira Basin, 18 = Agadir and Tarfaya Basins, 19 = Aaiun Basin, 20 = MSGBC (Mauritania-
Senegal-Gambia-Bissau-Conarky) Basin, 21 = Greater Niger Delta area (3 squares), 22 = Astrid Thrust Belt, 23 = Lower Congo Basin, 24 = Kwanza Basin, 25 = Namibe Basin,
26 = Orange Basin, 27 = Majunga Basin, 28 = Rovuma Basin, 29 Kenya, 30 = Somalia, 31 = deep shale detachment Nile Delta, 32 = Messinian salt DWFTB deepwater Niger Delta-
Levant Basin, 33 = Cyprus arc accretionary fold belt, 34 = Adana-Cilicia and Iskenderun-Latakia basins, 35 = Black Sea, 36 = South Caspian Sea, 37 = Makran accretionary prism,
38 = Krishna-Godavari Basin, 39 = Indo-Burma Ranges, 40 = West Luconia Delta, 41 = NW Borneo (Brunei, Sabah), 42 = Sandakan Basin, 43 = Mahakam Delta, Makassar Straits,
44 = East Sulawesi-Makassar Straits, 45 = Banggai-Sula, 46 = Molucca Sea Collision Complex, 47 = Cenderawasih Basin, 48 = Seram, 49 = Timor, 50 = Sumatra-Java accretionary
prism, 51 = Bight Basin, 52 = Hikurang accretionary prism. DEM from http://www.ngdc.noaa.gov/mgg/image/2minrelief.html.

the sedimentary section above a basal detachment (e.g. passive margins), used in the classification, and the nature of the basal detachment
or in response to stresses that affect much of the crust or the entire zones are discussed in this section in order to justify the new DWFTB
lithosphere, such as those found in accretionary prisms and collision belts classification presented herein and in Fig. 2.
(e.g. Rowan et al., 2004; Hamilton and De Vera, 2009). Consequently
DWFTBs have been subdivided into active and passive margin settings 2.2. Stress-system terminology
(Hamilton and De Vera, 2009; Krueger and Gilbert, 2009). Further
subdivisions are made on the basis of whether the basal detachments to Within the classification scheme proposed herein, the separation
the DWFTBs are developed within salt or shale lithologies. into gravity-driven and lithospheric stress driven systems seems a
There are three main problems with the existing classification reasonable, simple subdivision. However, gravity plays an important
schemes: 1) inaccurate terminology, 2) failure to encompass all types role in driving lithospheric plates and can also drive flow of hot,
of DWFTBs within the classification, and 3) gravity and lithospheric ductile crust in some regions (e.g. Tibetan Plateau, see review in
stresses driving mechanisms are not exclusively confined to passive Ghosh et al., 2006; SE Asia, Hall, in press). Consequently, use of the
and active margins respectively. These problems, the terminology term gravity-driven for just the systems confined to the sedimentary

Type 1 Type 2
Predominantly/ Mixed near field Predominantly/
Stress type exclusively and far field stress exclusively far field
near field stress stress
Type 2a Type 2b
Potentially any type of Continental Continental Accretionary
setting with a slope convergence convergence prisms
Tectonic setting to deepwater. In zones zones
practice predominantly
passive margins Type 2bi Type 2bii
Type 1a Type 1b
Detachment type Shale Salt Shale Shale Shale

Detachment dip O L O L L L

Landward = L, Oceanward = O

Fig. 2. Classification scheme for modern deepwater fold and thrust belts used in this review. For ancient type 2a and 2bi DWFTBs the basal detachment can be located in shale or in
salt, however no modern examples of salt detached type 2a and 2bi DWFTBs are known.
44 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

section could be accused of inaccuracy. Instead, where the potential zone illustrates there is not necessarily a simple distinction between
energy for driving deformation arises from uplift or sediment loading gravity-driven passive margin DWFTBs and tectonically driven
and results in gravity-driven deformation confined to deformation DWFTBs on active margins (see Section 4.2.1.1. for details). Passive
above one or more detachments that reside entirely within the margins are not synonymous with near-field stress driven deforma-
sedimentary section, we use the term near-field stress systems. Far- tion, and the same applies to active and collisional margins with
field stress is used for thin-skinned DWFTBs driven by lithospheric respect to far-field stresses. It should also be noted there is potential
stresses, and/or where gravity drives deformation in parts of the for far field stresses to cause deformation (e.g. inversion structures)
middle or lower crust that are sufficiently hot and weak. on Atlantic-type passive margins (e.g. Cloetingh et al., 2008), and to
cause folding of oceanic crust (e.g. the Central Indian Ocean fold belt,
2.3. Are near-field stress systems confined to passive margins? Beekman et al., 1996).

The prerequisites for near-field stress deformation are a basin with a 2.4. Tectonic setting: active margins and convergent zones
consistent regional slope into a deepwater area, a weak detachment
zone, and a trigger of some kind (usually uplift or delta progradation). The term active margin covers either convergent or Pacific-type
This set of conditions is not exclusive to passive margins, although the margins associated with subduction of oceanic crust, or transform
greatest number of near-field stress DWFTBs does occur on passive margins, but not continent–continent, continent–arc collision, or
margins (e.g. Tari et al., 2003; Rowan et al., 2004). For example the intracontinental convergence settings (e.g. Condie, 1997). Unfortu-
collisional setting of the Mediterranean produced the Messinian nately all these settings are lumped together within the term active
evaporates (Cita, 1983), which also acted as detachment zones to margin in the classifications of Hamilton and De Vera (2009) and
DWFTBs, e.g. NE Mediterranean Cyprus Arc (Adana-Cilicia and Krueger and Gilbert (2009). In this classification far-field stress driven
Iskenderun-Latakia basins; Bridge et al., 2005), and deepwater Nile DWFTBs are subdivided on the basis of tectonic setting into active
Delta (Loncke et al. 2006). The Sandakan Delta (Fig. 3) developed when margins (i.e. subduction zones and transform margins) and conti-
uplift of Borneo triggered sedimentation and forced progradation of the nental convergent zones.
delta into the southern axis of the developing Celebes Sea spreading Continental convergent zones encompass deep water fold and thrust
centre (Fig. 3; Balaguru and Hall, 2009). The crust of northern Borneo is belts that formed during continent–arc collision or the early-stage of
composed of uplifted deepwater sediments, overthrust by ophiolites continent–continent collision, and in regions where a complete Wilson
that probably overlie oceanic or forearc crust (e.g. Hutchison et al. 2000). Cycle has failed to occur, e.g. thinned crust in an intraplate setting
Consequently, in terms of both the crustal type, and tectonic setting the marginal to an orogenic belt (intracontinental convergence zones).
Sandakan Delta did not form on a typical passive margin. SE Asia displays a wide range of DWFTB settings, including those
One of the success stories of hydrocarbon exploration in DWFTBs is related to different stages of subduction and early-stage continent–
the NW Borneo margin where up to 10 km thickness of deltaic continent or continent–arc collision, and obduction (e.g. NW Borneo
sediments accumulated during the Neogene (e.g. Sandal, 1996; Hall and the Banda Arc, Fig. 3; Searle and Stevens, 1984; Hall and Wilson,
and Nichols, 2002; Ingram et al., 2004; Morley and Back 2008). 2000), as well as intracontinental convergent zones. The Makassar
Shortening in the DWFTB was driven both by near field stresses with a Straits between Borneo and Sulawesi are an example of an
linked growth fault system on the shelf, and by far-field stresses intracontinental convergence zone, where a deepwater trough
associated with regional shortening (e.g. Morley et al., 2003; Hesse containing a DWFTB on the west side of Sulawesi has developed
et al., 2008; Morley et al., 2008; King et al., 2009). This mixture of over thinned, rifted continental crust (Figs. 4 and 5; Hall et al., 2009).
shallow gravity, and deep-seated stresses in an early-stage collision Intracontinental convergence zones have no place in classification

Fig. 3. Distribution of deepwater fold and thrust belts in SE Asia, oblique DEM image.
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 45

schemes where the subdivision is into active and passive margins 2) DWFTBs partially or completely driven by far-field stresses, with
(Rowan et al., 2004; Hamilton and De Vera, 2009; Krueger and Gilbert, a basal shale detachment that typically dips landwards.
2009).
2ai) Linked extensional–compressional systems in zones of
2.5. Classification scheme inter- or intra-continent convergence (gravity sliding/
differential loading/far-field lithospheric stresses), e.g.
On the basis of the tectonic and structural divisions discussed NW Borneo margin; Columbus Basin, Trinidad.
above, DWFTBs are be divided into the following categories (Fig. 2): 2aii) Weakly linked or unlinked extensional–compressional
structures in DWFTB systems on early stage zones of
1) Near-field stress-driven linked extensional-compressional sys- inter- or intra-continent convergence, mostly or entirely
tems (mostly, but not exclusively found on passive margins), driven by far-field lithospheric stresses, e.g. East Makassar
commonly related to gravitational instability of a seaward Straits, Indonesia; Banda Arc, Indonesia.
prograding sedimentary wedge. 2b) Weakly linked or unlinked extensional–compressional
structures in DWFTB systems on active margins (sub-
1a) Predominantly offshore dipping, basal overpressured duction and transpressional margins), mostly or entirely
shale detachment (e.g. Orange Basin, South Africa; Bight driven by lithospheric stresses, e.g. Barbados accretionary
Basin, S. Australia; Niger Delta, Nigeria – mixed offshore prism; Makran, Iran and Pakistan, Indo-Burma Ranges,
and landward-dipping detachment). Bangladesh and Myanmar, South Caspian Sea.
1b) Predominantly offshore-dipping salt detachment (e.g.
West African margins of Gabon, Congo, Angola; Gulf of These categories and associated hydrocarbon provinces are
Mexico, Brazil Santos Basin). discussed in the following sections.
114°

116°

118°
110°

112°

NW Borneo Trough
0m Sandakan
Fold and Thrust belt 20
Delta
Rajang/West
m

Luconian Delta
00


10

Luconia m
m 00
Platform 200 10


Tarakan
Delta


Borneo

100 km
s

Shale mini-basin province 0°


ait

offshore Sabah
Str

Mahakam
ar

Delta
ss

Area containing deepwater


ka

fold and thrust belt


Ma

Normal faults 2°

West
F
Fold axes Sulawesi
Fold belt
Seafloor
Thrust faults
bathymetry

Fig. 4. Regional map of Borneo showing the location of DWFTBs around the island.
46 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

W Offshore Onshore E
Makassar Straits Deepwater fold West Sulawesi Ophiolite
Seafloor and thrust belt Inverted rift basins 0

Thinned continental crust

Mantle

40
Cenozoic sedimentary rocks

W Break thrusts and detachment folds E


Detachment folds 2

Sea floor 3

6
Broad detachment 20 km
7

Fig. 5. Regional cross-section through the Makassar Straits based on Calvert and Hall (2007). The West Sulawesi DWFTB is developed on thinned continental crust of the Makassar
Straits.

3. The characteristics of salt and shale detachments translation of sediments overlying the basal detachment (e.g. Cobbold
and Szatmari, 1991; Duval et al., 1992; Maudit et al., 1997; Rowan
3.1. Gravity sliding and gravity spreading et al., 2004). Loading by large deltas like the Niger, which create
depocenters over 10 km thick, can cause a large area of the basal
Gravity gliding (or sliding) and gravity spreading are the two basic detachment to rotate during deposition and dip towards the continent
types of near-field stress driven deformation of a sedimentary wedge (at least under the DWFTB, diapir and outer growth fault provinces;
that progrades into deeper water. Gravity gliding occurs by the rigid Worrall and Snelson, 1989; Morley and Guerin, 1996). More localized
translation of a rock mass down a slope, while gravity spreading is the variations in loading can initiate flow in a thick mobile unit from
flattening and lateral spreading of a rock mass under its own weight regions of high pressure to low pressure, without the need for a
(Ramberg, 1981; DeJong and Scholten, 1973; Fig. 6). Gravity surface or basal slope (e.g. Kehle, 1989; Fig. 6B). Gravity spreading
spreading rarely affects the complete sedimentary section and is becomes a more important driving mechanism where a thick mobile
often limited to a thick mobile zone (usually either overpressured, unit is present (e.g. Morley and Guerin, 1996; Vendeville, 2005), but
undercompacted muds or salt) at the base of the gravity-driven in such circumstances gravity sliding and gravity spreading tend to
system. In contrast to the original definition, gravity spreading of a operate in tandem (Brun and Merle, 1985).
mobile zone in DWFTBs would occur in response loading by overlying
sediments, not just under it's own weight. 3.2. Salt detachment zones
Gravity gliding is usually associated with linkage of up-dip
extension with a down-dip contractional toe region via a detachment DWFTBs are either associated with shale or salt detachment zones.
zone and covers a wide range of temporal and spatial scales. There are Important differences in structural style arise from these different
the geologically instantaneous, shallow slides associated with mass detachment types and consequently classification schemes intro-
wasting that can range in size from small landslips covering a few duced subdivisions on this basis (Rowan et al., 2004; Krueger and
hundred square meters in area, to giant submarine slides that can Gilbert, 2009), which are also used in this classification (Fig. 2).
cover over 100 km2 (e.g. Embly, 1976, Canals et al., 2004; Gee et al., Although the typical structural components of salt and shale
2007). Such slides are often triggered by short-term events e.g. detachment systems are similar (up-dip growth faults, diapirs,
storms, earthquakes and high rainfall. Of greater interest to this compressional toe region), in detail the structural styles, timing of
review are the gravity gliding systems on passive margins where the deformation, location of deformation, and trigger mechanism can be
detachment is buried deeper (~1 km or greater), extends for tens of very different. Fig. 7 is a schematic illustration of two similar
kilometers in the transport direction and develops as a result of long- structures, one with a mobile shale unit the other with a salt
term geological processes (e.g. high sedimentation rates, uplift of the detachment, to illustrate some of the differences between the types
adjacent continental area resulting in tilting of the margin). Gravity of mobile unit. Highly overpressured saline fluids can be encountered
gliding can occur along a sharp, narrow detachment surface or along a immediately below a salt intrusion and in a brine halo surrounding the
zone of ductile deformation (where gravity spreading may also salt as warm, overpressured fluids rising from depth dissolved salt and
operate). followed fault and fracture permeability adjacent to salt masses and
Passive margins commonly develop extensive oceanward-inclined either became trapped in sealed compartments or migrated upwards
detachment surfaces that facilitate oceanward shallow gravity-driven towards the surface (Warren, 2006). Conversely overpressured shale
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 47

A Top of sediments prior to Extensional domain


gravity sliding

Compressional
‘toe’ domain

Narrow detachment zone

B Diapir Extensional domain


Compressional
‘toe’ domain

2 1

C Extensional domain
Compressional
‘toe’ domain Diapirs

Pre-deformation top of mobile detachment zone

Brittle sedimentary rocks within gravity-driven


deformation belt
Mobile detachment zone (viscous or plastic
material behaviour)
Fig. 6. Schematic illustration of gravity spreading and gravity gliding mechanism. A) Gravity gliding down an inclined slope on a thin detachment, B) Sediment loading causing
gravity spreading employing a thick detachment zone on a flat detachment surface, C) mixed gravity gliding and gravity spreading on a thick detachment zone with both offshore
and onshore inclined detachment surfaces.

sequences tend to show a downwards increase in overpressure over a subsidence, regional uplift to form gentle slopes or differential loading
broad interval typical of disequilbrium compaction. Shale diapirs are by sediment can trigger deformation. (e.g. Kehle, 1989; Worrall and
highly overpressured and will rise by hydraulically fracturing into the Snelson, 1989; Duval et al., 1992; Spathopoulos, 1996, Marton et al.,
overlying strata, and stoping blocks of country rock (Morley, 2003a,b). 2000).
As discussed by Morley and Guerin (1996) mobile shales tend to In collision belts salt is typically present within the platform and/
deactivate with time due to dewatering and loss of overpressure, or or foreland basin sequences (e.g. Idaho–Wyoming segment of the
may change with time from disequilbrium compaction overpressures Rocky Mountains; Romanian Carpathians; Jura Mountains, Alps;
to compacted shales fractured by inflationary overpressured fluids Zagros Mountains; Siwaliks, Pakistan; Oman Mountains, see review
Conversely the intrinsic rock properties of salt means that it can in Warren, 2006). Salts form, apart from shales or even porous
deform, flow, migrate through the sedimentary section, and still sandstones, important detachment horizons in the external zones of
maintain its ductile nature, and re-activate even when long time fold and thrust belts, but generally salt is not reported as playing an
periods have elapsed between deformation events (see review in important role in deformation of the more internal zones. As the
Warren, 2006). South Atlantic margin demonstrates, salt can extend from onshore
Actively subsiding basins with restricted communications to the into the deepwater on passive margins, but it is questionable
open seas are required for the development of large salt deposits. whether salt can survive long enough to act as a regional
Seepage of salt water across a barrier that partly separates the sea detachment horizon during collision of an accretionary wedge
from a restricted evaporite basin can lead to extensive deposition of with a passive margin, or whether it has been mostly extruded
evaporites (Warren, 2006). Two of the main occurrences of such and recycled in the deepwater sequences prior to collision or during
conditions are during the late syn-rift stage (e.g. Central Atlantic), and its earliest stages. An example of extensional salt-related raft
early rift to drift transition on passive margins (e.g. South Atlantic tectonics along the Cotiella thrust sheet in the Southern Pyrenees
margins of Gabon and Angola and Brazilian Santos and Campos has been recognized, but this thrust reactivates a salt weld (McClay
basins). Hence salt detachments associated with DWFTBs are most et al., 2004). In this context the example of the Moroccan Rif is
common on passive margins (the collisional setting of the Messinian discussed below.
evaporite in the Mediterranean is an important exception). The west The onshore Prerif Thrust Sheet of the Moroccan Rif is an example
African–east South American margin is a classic example of a passive of the deformation style associated with a passive margin to
margin, post-rift salt setting (e.g. Guardado et al., 1989; Teisserenc deepwater section that has been extensively affected by salt diapirs
and Villemin, 1989; Duval et al., 1992; Spathopoulos, 1996; Tari et al., and nappes (Suter, 1980a,b; Wildi, 1983; Chalouan et al., 2008). The
2003. Salt is very weak, and seaward tilting during post-rift thermal rocks of the Prerif Thrust Sheet lay in a passive margin setting until
48 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

A Mud volcano
Mud chamber
Zone of overpressure

Fluid flux

Overpressured, Weld, where mobile


but not mobile shales shale was expelled
Compacted, previously Undercompacted, Stoped blocks
Migration and concentration
overpressured shale overpressured within mobile
of overpressured fluids
Sediments with gas effects and shales shales
inflationary overpressures , which Shale intrusions Ductile folds in
Poorly imaged bedding
reduce reflectivity on seismic data mobile shale

B Zone of overpressure higher beneath


salt canopy

Weld, where salt


Fault zone or gumbo, commonly overpressured was expelled

Salt Brine halo from salt dissolution

Fig. 7. Schematic comparison of structural style and overpressure distribution associated with thick mobile shale and mobile salt detachments. A) Mobile Shale, B) Salt. Partly based
on Warren et al. (in press) and observations from the Niger Delta and Brunei Darussalam (Morley and Guerin, 1996; Van Rensbergen et al., 1999; Morley, 2003a,b).

collision of the N African margin with the Intrarif continental margin when subject to regional compression expelled the salt and
fragment and the Alboran Domain induced shortening during the further broke into highly discontinuous imbricates. The Prerif example
Oligocene and Neogene (Michard et al., 2007). This thick, extensive, together with the Permo-Triassic Haselgebirge Formation in the
poorly organized Prerif Thrust Sheet is several hundred kilometers Northern Calcareous Alps (Spötl, 1989; Mandl, 1999) illustrates that
long and 30–50 km wide. It is composed of large and small oval to thick depositional salt masses in a passive margin setting may be too
linear blocks of rock that range in size up to entire hillsides several deformed by gravity-driven flow to form a continuous detachment
kilometers long that are arranged either chaotically, or in a loosely horizon during later collision. Such a margin will form a poorly organized
imbricated pattern within a fine grained matrix of Cenozoic tectonic unit during continent–continent collision. However, evidence
mudstone. Blocks of Triassic salt up to 100's m across are extensively from other ancient DWFTBs indicates that deeper water flysch deposits
present. The unit has been interpreted as an olistostrome or tectonic in Type 2 thrust belts can be associated with evaporite detachments,
melange, with most blocks consisting of Mesozoic shelf-slope examples include Albania (Velaj, 2001; Roure et al., 2004) (Permo-
sediments; deepwater, trench associated lithologies are generally Triassic platform sequence), Carpathians (Burdigalian & Badenian
lacking (Wildi, 1983; Chalouan et al., 2008). Cenozoic turbidites have evaporites in foreland basin sequence) , and the Apennines (Messinian
been thrust over the Prerif Thrust Sheet (Suter, 1980a,b). The geology evaporites in foreland basin sequence) (e.g. Letouzey et al., 1995).
of the Prerift Thrust Sheet is best explained as a highly deformed For the collisional and active margin DWFTB settings discussed
halokinetic passive margin. Such a pre-weakened and deformed passive here overpressured shale is considered as the primary mobile unit.
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 49

3.3. Shale detachments supporting some of the overburden or tectonic stresses and becoming
overpressured as a result (e.g. Osborne and Swarbrick, 1998).
Shale detachments in thrust belts fall into four main categories: 1) Undercompaction of shales overpressured by disequilbrium compac-
compacted shales that act as weak, easy slip horizons due to inherent tion has been detected from 3D seismic and well data in the
material weakness (in relation to adjacent units) without large developing decollement of the Nankai Trough (Bangs and Gulick,
overpressures, 2) compacted shales with large overpressures, 3) thin 2005).
(meters to tends of meters thickness) undercompacted shales with Shale detachments in DWFTBs are widely associated with high
high overpressures, and 4) thick (100's meters–kilometers thickness) pore fluid pressures that result from disequilbrium compaction,
undercompacted, overpressured shales capable of large-scale viscous inflation or a combination of the two (e.g. Dahlen, 1990; Fisher and
flow. Hounslow, 1990; Moore et al., 1995; Fisher et al., 1996; Fisher and
Undercompacted shale detachment zones require preservation of Zwart, 1997; Moore and Tobin, 1997; Screaton et al., 1997; Bilotti
porosity during burial prior to DWFTB formation. Such preservation and Shaw, 2005; Morley, 2007a). Indirect evidence of widespread
arises due to disequilibrium compaction, in which the rate of fluid overpressured conditions, associated with shale detachments, is
expulsion (due to compaction) in a formation is slower than the rate provided by the numerous mud volcanoes and fluid escape features
of vertical or tectonic loading, and results in formation fluids present in most DWFTBs (Fig. 8B). The deep-seated nature of these

Water saturated incoming


trench sediments
A Mocene-Recent
slope basin fill HIKURANGI TROUGH
Shelf
0
Protothrusts

3
km

6 50°C

100°C PACIFIC PLATE


9 Decollement 10 km
150°C Subducting pelagic sequence Subduction rate 42 mm/yr VE ~ 2
12

Brittle deformation, dilation, common extension veins, scaly shear fabrics Protothrusts, cataclasis, brittle-ductile shear

Smectite to illite phase transition Porosity reduction due to compaction and tectonic deformation
100-150° C drives fluids from sediments at low tempertaure and pressure
(<5 km, ~10-20% ~60-70%)
Clay mineral dehydration and hydrocarbon
generation. (> 80% fluid accreted is released)
(3 m 3 m yr -1 H 2 0 added to fluid system)
Imbricated foundation of pre-subduction
Fluid migration pathway Base of gas hydrates Late Cenozoic accretionary wedge
Cretaceous-Palaeogene rocks

1 = Mud pipe to surface 7 = Anastamosing channel


B mud volcano field system feeding fan
2 = BSR, base of gas hydrates 8 = Piggy-back basin in syncline
3 = shallow fluid escape pipes 9 = depression associated with
9 and pockmarks at surface mud volcano field
9 4 = Crestal normal faults 10 = channel accommodating
7 5 = Saddle area in folds down-slope flow of mud
5 6 = Deepwater fan from mud volcanoes
6
11 = slump scar from landsides on
forelimb of anticline
12 = Submarine slide MTC
10 13 = Basal detachment

11
9 3
12
2 Stratigraphy permeability
0
4 8 1 Fluid movement along faults
8
Free gas
Gas hydrate

3 km
13

Simple anticline with Mature anticline with landslides and


crestal normal faults and degradation complex in forelimb
break thrust on forelimb

Fig. 8. Illustrations of fluid flow in deepwater fold and thrust belts associated with shale detachments. A) Cross section through the Hikurang Subduction zone cross section (Barnes
et al., 2010). B) Schematic 3D section showing typical features associated with growing deepwater folds (based on Morley, 2009a and Barnes et al. 2010).
50 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

features is shown by seismic reflection data that reveals fluid pipes considered to be typical of low-temperature thrust fault zones
rising from large thrust faults (e.g. Oregon margin, MacKay et al., dominated by fracturing (Hung et al., 2007). According to Hung et
1992; Niger Delta, Cobbold et al., 2009; Baram Delta Province, al. (2007) the physical properties of the Chelungpu Fault zone with
Morley, 2009a; Hikurang subduction zone, Barnes et al., 2010; respect to adjacent host rocks are: relatively low resisitivity (breccia
Fig. 8) and sea floor sampling of mud volcanoes which shows many zone about 40% lower) and density (20–25% reduction), and high
have high heat flows and the fluids contain thermogenic hydro- Poisson's ratio (~0.4). The low resistivity is due to infiltration of
carbons (e.g. Deville et al., 2003; Dolan et al., 2004; Deville et al., drilling mud into highly fractured breccia, while the low density and
2006; Zielinski et al., 2007; Warren et al., in press; Fig. 9). High pore high Poission's ratio in the gouge are related to high clay content and
fluid pressures are the primary explanation for the very low or fluid. The gouge shows relatively high concentrations illite or
detachment strength required to generate the low critical taper of mixed layer smectite/illite compared with other inactive fault zones in
DWFTB wedges (e.g. Dahlen, 1990; Bilotti and Shaw, 2005; Morley, the area, suggesting conversion of smectite to these minerals caused
2007a). by (minor) frictional heating during the Chi-Chi earthquake (Hung
The presence of low friction coefficient (below 0.25) clay minerals et al., 2007).
also plays an important role in shale detachment weakness (e.g. Kopf Shale detachments commonly have high permeability related to
and Brown, 2003), and the association of DWFTBs with high pore fluid fractures, or conditions favoring disequilbrium compaction, which
pressures is not universal. For example Suppe (2007), Hung et al. enables the presence and flow of overpressured fluids (e.g. Mann and
(2007) and Yue (2007) describe the weak Chinshui Shale detachment Kukowski, 1998). Shale detachments contain numerous sources of
in the Chenglungpu Fault, from Taiwan where the surrounding rocks permeability heterogeneity related to clay mineralogy, variations in
and the fault zone exhibit only hydrostatic pore fluid pressures. sedimentary unit type, diagenesis and porosity distribution. For
However, the Chenglungpu Fault links to the basal detachment in example radiolarians can cause high initial porosities that abruptly
Taiwan, which probably is associated with high pore fluid pressures collapse during burial, while turbidite sands encased in shales can
(Yue, 2007), and a transient pulse of high overpressure during faulting compartmentalize fluid flow until they cement up (Underwood,
cannot be eliminated as a possibility (Hung et al., 2007). The 2007). More lithified detachments may generate a permeable dilatant
Chenglungpu and adjacent Sanyi fault zones offers a rare insight breccia (Maltman et al., 1993).
into one type of shale thrust zone, since two wells were drilled by the Fluid is typically pumped along a detachment immediately after it
Taiwan Chenglungpu-fault Drilling Project, which obtained continu- fails (Maltman et al., 1997). Shearing of the detachment and loss of
ous core across the fault zones (depths between 1013 and 1710 m). overpressured fluids will tend to reduce porosity in relation to
The Chenglungpu Fault was activated during the Chi-Chi earthquake adjacent sediments beyond the shear zone (Taylor et al. 1990; Brown
(Hung et al., 2007). As described by Sone et al. (2007), the main et al., 1994). The extensional shear and hydrofracture networks
Chenglungpu Fault zone is about 200 m wide, and displays a observed in many accretionary prisms require that pore fluid pressure
transition from a broad zone of weakly to intensely fractured rock, is at about lithostatic pressure in order for fractures to remain open
three narrow zones (about 3–4 m wide) of breccia, foliated breccia (Berhmann, 1991). These fractures will tend to be low-angled since
and fault gouge form the principal displacement zones. Black, fine- the minimum principal is vertical (Brown et al., 1994). A high fluid-
grained, slicken-lined layers less than 10 cm thick within the principal pressure, dilatant zone of high porosity about 14 m thick characterizes
displacement zones may represent the products of focused, intense the basal decollement in the northern Barbados Ridge (Shipley et al.,
co-seismic shearing. The paucity of veins within the fault zone is 1994). An abrupt transition from lateral compression within the

Fig. 9. Results of bottom sampling of the deepwater area of Brunei. A) Heat flow, B) Hydrocarbon content from piston cores. Data from Zielinski et al. (2007). Data points are overlain
on a map of the waterbottom based on 3D seismic data, showing the locations of anticlines (from Morley, 2007a).
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 51

Table 1
Summary of the key characteristics of shale detachments in DWFTBs. 1 = Morley and Guerin (1996), 2 = Cobbold et al. (2009), 3 = Colten-Bradley (1987)), 4 = MacKay et al. (1992),
5 = MacKay (1995), 6 = Cobbold et al. (2004), 7 = Underwood (2007).

Type 1a Types 2a, 2bi Type 2bii

Origin of detachment weakness Overpressure, weak, poorly lithified mudstones, Overpressure, poorly lithified mudstones, Compacted shale with weak clay
or compacted shale with weak clay minerals or compacted shale with weak clay minerals minerals and overpressure (3–7)
Origin of overpressure Disequilbrium compaction, inflationary Inflationary due to burial diagenesis during Inflationary, mostly from fluids
overpressures due to diagenesis during burial and underthrusting of sediments derived from sediments subducted
burial (1, 2) beneath basal detachments in the lower plate (7)

Barbados accretionary wedge, to vertical compaction within the upper them by means of disequilibrium compaction (e.g. Morgan et al.,
part of the detachment zone has been identified from a magnetic 1968; Evamy et al., 1978; Knox and Omatsola, 1989; James, 1984;
susceptibility survey (Housen et al., 1996), the transition is marked by Osborne and Swarbrick, 1998; see review in Tingay et al., 2009).
high pore fluid pressures (Shipley et al., 1995), that were developed in Subsequently inflationary overpressures internal to the shale body
conditions of brittle rather than ductile deformation (Housen, 1997). may develop (e.g. hydrocarbon cracking, tectonic compaction,
The Amazon Fan provides an example of an overpressured, compacted smectite–illite transition (Osborne and Swarbrick, 1998; Tingay et
shale detachment on a passive margin (Cobbold et al., 2004). al., 2009). Such overpressure mechanisms have the potential to
The causes of long-term fluid fluxes into the basal detachment preserve porosity within the shale mass and may promote a more
zone vary depending upon tectonic setting and these variations can ductile type of detachment than is generally observed in accretionary
influence deformation style and history (Table 1). There is a prisms (Ings and Beaumont, 2010). One possible example comes
potentially significant difference between the detachments in parts from eastern Bangladesh, where a mixed disequilibrium compaction
of large deltas and accretionary prisms, related to sedimentation rates. and inflationary origin for the overpressures across a 1000 m thick
In typical accretionary prism settings deposition rates are low, and interval has been determined (Zahid and Uddin, 2005; Hossain et al.,
deformation can occur over many tens of millions of years. 2009). Such interpretations are controversial, and whether broad
Consequently detachments tend to occur in fractured, overpressured zones (i.e. 100's m thick) of ductile, ‘mobile’ overpressured shales
compacted shales, fed by overpressured fluids originating from the can exist at depth is discussed separately in the following section
down going slab. Variations in inflationary overpressure production in (Section 3.4).
the down going slab can affect the structural style. For example in the
Cascadian accretionary prism changes in vergence are accompanied 3.4. Do thick mobile shale zones exist?
by differences in the detachment level from 400 m to 1200 m above
the igneous basement, and oblique and strike-slip faults are present at The presence of thick mobile shales in a deltaic setting up to
the abrupt changes in vergence (MacKay et al., 1992; MacKay, 1995). several kilometers thick, buried between depths of 3 and 10 km, and
In the oceanward-verging zone sediments in the down going slab composed of undercompacted, overpressured shales that deform in
reach a maximum temperature range of 120°–145 °C, while in the highly ductile manner has been widely accepted in scientific literature
landwards-verging zone maximum temperatures reach 135°–160 °C (e.g. Evamy et al. 1978; Morley and Guerin, 1996; Sandal, 1996;
(Underwood, 2007). The additional 15°–20 °C temperature under the McClay et al., 1998; Ajakaiye and Bally, 2002). However, the existence
landwards-verging zone probably caused significant extra compac- of thick, deeply buried mobile shales are, at present, a controversial
tion and permability loss associated with the smectite–illite transi- aspect of delta tectonics due to improved seismic resolution, and new
tion, which consequently favoured the development of relatively high observations about the extent of chemical diagenesis in shales. For
disequilbrium compaction overpressures. As summarized in Table 2, example Ruarri et al. (2009) concluded that shale movement is not
Moore et al. (2007) provide an excellent summary of the changes in akin to salt, and that: 1) deeply sourced shale diapirs are highly
detachment behaviour found in accretionary prisms with respect to unlikely, 2) increasing seismic resolution will eventually demonstrate
changing temperature and pressure. thick mobile shale masses do not exist, and 3) lateral flowage of shale
In deltas associated with high sedimentation rates, detachments is highly unlikely. Investigation of shale diagenesis from Upper
are developed within the pre-delta or earliest delta-stage marine Cretaceous mudstones, offshore Norway by Thyberg et al. (in press)
shales. Due to their rapid burial, overpressures are first developed in demonstrated the smectite–illite transition releases fine-grained

Table 2
Summary of the observed cements, veins and relevant structural fabrics of shale detachment zones in accretionary prisms at various temperature and metamorphic facies ranges.
From Moore et al. (2007).

Temperature and pressure range Observations and source localities

b 100 °C Unmetamorphosed No veins and cements through nearly all cored rocks in ODP. Minor carbonate in Barbados drill cores. Extremely rare quartz,
carbonate and sulphides in Nankai Trough. Accreted terrigenous deposits on Barbados Island lack any carbonte/quartz veins or
cements. Faulting and stratal disruption includes cataclasis of sand sized particles with widespread development of scaly fabric
in mudstones
100°–150 °C; 5 km Typical Zeolite Facies Evidence for cementation and veining variable. Locally significant carbonate cementation along fault zones. Quartz cementation/
veins rare and minor. Cataclasis of sand–sized particles widespread. Development of scaly fabric in mudstones. Incipient
pressure solution cleavage.
150°–300 °C; 5–15 km Typical Prehnite Quartz- and carbonate-veining common. Cataclasis in sands and veined materials temporally alternates with pressure solution.
Pumpellyite Facies Cleavages common, especially visible in more coherent units. Quartz veins/cements common in the 200°–300 °C ranges. Rare
pseudotachylyte. Veining and cementation facilitate change from velocity-strengthening to a velocity weakening, earthquake-
prone rheology.
150°–350 °C; 15–27 km Classic HP-LT Often coherent layering with multiple fold generations, local stratal disruption predating metamorphism. Pressure solution
(e.g. Franciscan Blueschists of eastern belt) common. Progressive replacing of brittle fabrics with more ductile fabrics
N 300 °C; 8–27 km Prehnite–Actinolite, Widespread foliation development, pressure solution, microscopically apparent recrystallisation during development of
and Blueschist–Greenschist Facies metamorphic assemblages. Many terranes surprisingly coherent, with multiple phases of folding. Brittle deformation during
peak metamorphism not obvious, but a previous history is suggested by the stratal disruption in some areas.
52 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

quartz that helps stiffen mudstones. 1–3 μm spherical, discrete grains diapirs diminishes, and diapirs are re-interpreted as shale-cored
or short chains and clusters of quartz form a pervasive network that anticlines. Present understanding of chemical compaction assumes
developed when burial reached depths around 2500 m or 80°–85 °C. there is no overpressure control on retardation. Conversely, the
Helset et al. (2002) have proposed that while overpressure retards evidence in favor of mobile shales suggests such diagenesis
mechanical compaction, it does not retard chemical compaction, and retardation mechanisms exist. For example, there is some evidence
consequently diagenetic processes may control the timing and for overpressures causing either a change in the type of diagenesis
magnitude of overpressure rather than disequilibrium compaction. (Xie et al. 2003), or retardation of diagenesis in sandstones (e.g.
Mud volcanoes apparently provide evidence for deep, under- Osborne and Swarbrick, 1999) and chalk (Hardman, 1982), particu-
compacted, overpressured shales. However, if they originated as larly if hydrocarbons are part of the overpressured fluids.
diatremes this may not be the case (Brown, 1990). Diatreme-type
mud pipes develop as a consequence of deep gas generated 4. Structural and petroleum system characteristics of the different
overpressures that rise upward in a hydraulically fracturing column types of DWFTB
and mix with higher undercompacted muds and fluids to generate
mudflows at a shallower level (Brown, 1990). However, the existence 4.1. Near-field stress-driven linked systems
of diatreme-type pipes does not mean that all mud pipes can be
explained in this way (Brown, 1990). In Brunei for example Middle 4.1.1. Introduction
Miocene shoreface clastics were intruded in two phases (pre-folding Near-field stress-driven DWFTB have been reviewed by Rowan et
and post-folding) by overpressured mudstones derived from the al. (2004) in considerable detail, although, as noted in Section 2.3 we
underlying Setap Shale Formation (Morley et al., 1998; Morley, have not limited this category exclusively to passive margins. Tari et
2003b). The Setap Shale Formation forms the basal ‘mobile’ shale zone al. (2003) proposed sub-divisions of salt-detachment passive margins
for the Baram Delta Province and in the subsurface overpressures types based on observations along the West African margin. They
increase downwards towards the Setap Shale Formation as expected noted that the extensive post-rift Aptian-age salt found offshore
for a classic disequilbrium compaction detachment zone. Furthermore Cameroon, Equatorial Guinea, Gabon, Congo and Angola was
the porosity-effective stress relationship follows the classic disequil- relatively thick and widespread, and comprised a system of up-dip
brium compaction model within the Setap Formation, not the extension, and down-dip contraction where salt nappes, canopies and
inflationary overpressure trend expected for a chemical compaction toe-thrusts and folds accommodate most of the shortening. In
model (Tingay et al., 2009). contrast the Late Triassic and Early Jurassic syn-rift salt in northern
The issue of whether thick mobile shales are present at depth West Africa (Morocco, Mauritania, Senegal, The Gambia and Guinea–
arises in regions of shale mini-basins and diapirs, and large growth Bissau) was isolated by syn-rift highs and did not necessarily link with
faults; the Niger Delta is particularly important to the debate. The up dip extension. Tari et al. (2003) identified the rate of sediment
presence of downbuilding synformal depocentres, and strong evi- loading and resulting slope steepness, and efficiency (i.e. thickness of
dence for pinching and swelling of the inferred mobile shale zone salt, salt detachment continuity) of the salt detachment as the first
from seismic reflection data led Morley and Guerin (1996) to order controls on the observed variations in salt deformation style
conclude that parts of the Niger Delta (particularly the outer shelf along the West African margin.
and slope) were underlain by a thick mobile shale unit, an The first order control scheme of Tari et al. (2003) was tailored
interpretation reconfirmed by Ajakaiye and Bally (2002). Numerical particularly for salt, and the differences in continuity of salt
modeling of shale-detachment structures like the Baram Deltaic detachment between salt deposited in a syn-rift and post-rift setting.
province and the Niger Delta indicates that a thick mobile shale Such distinctions are not exactly applicable to shale detachments,
section is required to generate the counter-regional fault province since continuous, thick shale horizons, with the potential to become
that lies in the outer shelf region (a thin detachment does not produce overpressured are present much more frequently than thick salt
counter-regional faults), and that a thin shale detachment is present layers, whilst overpressures in shales are transitory and bleed off over
in the outer fold and thrust belt (Ings and Beaumont, 2010). A study time (unlike salts that are always mechanically weak at typical
by ExxonMobile of the inner fold and thrust belt and diapir belt of the subsurface temperatures and pressures). Nonetheless the general
Niger Delta utilized seismic reflection geometries, velocities and approach of the Tari et al. (2003) classification is also appropriate for
gravity data to conclude that the seismically transparent ‘mobile’ shale detachments (Fig. 10).
shale zone was a thick, low density, low velocity, overpressured, There does not appear to be any clear-cut variation in basic
ductile zone (Wiener et al., 2009). This zone corresponds with the characteristics, such as fold wavelength, between the different
region of thick mobile shales required by the numerical modeling of detachment lithologies (Fig. 11). The clearest control on fold
Ings and Beaumont (2010). wavelength, amplitude and spacing is related to the maximum
A distinction needs to be made between whether thick shale stratigraphic thickness of units overlying large-scale shale detach-
sequences, capable of viscous flow exist, and whether shale diapirs ments, which varies from about 1 km to ten or more kilometers
similar to large salt-like diapirs can form (e.g. Van Rensbergen et al., (Fig. 12). The use of stratigraphic thickness is a crude parameter, but is
1999; Morley, 2003a). Possibly there are no shale equivalents to the most straightforward one that can be applied to all DWFTBs, and
classic salt diapirs, instead shale ‘diapirs’ are either narrow pipe like has been identified as an important parameter in numerical modeling
intrusions, or the larger ‘diapirs’ are actually shale-cored anticlinal of folding vs. faulting behaviour above detachments (Simpson, 2009).
folds, related to buckling and compression, not active diapiric Ideally the thickness of the dominant competent member should be
processes (e.g. Duerto and McClay, 2002). This folding origin for used. But in many DWFTBs the generally fine grained, poorly lithified
diapir mini-basin provinces may be applicable to both the Niger Delta nature of the section suggests that the presence of strong mechanical
(Wiener et al., 2009), and NW Sabah (King et al., 2009). contrasts within the stratigraphy are unlikely, or will be only locally
In summary there remains a strong case for the presence of thick present. Elastic strength will gradually increase with depth due to
mobile shales in parts of large deltas, although undoubtedly thin compaction and diagenesis, which is why stratigraphic thickness
detachment zones are present in many deltas and DWFTBs. Thick, shows a correlation with fold wavelength. The presence of controlling
ductile mobile shales may have been too widely applied in the past, competent layers will be more important on passive margins affected
but equally it is premature to completely eliminate them from all by salt, where the age range of units involved in the DWFTB is greater,
interpretations (see the Timor example, Section 4.3.1). Ductile mobile and higher degrees of lithification and compaction will have
shale detachments may still be present even if evidence for shale developed more mechanical contrasts within the stratigraphy. Larger
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 53

SW NE
0
Atlantic
Depth (km)

Top Pliocene
5 Top Miocene
Top Cretaceous
A
Top Cenomanian
10
0 40 km
Base Rift

W E
0
Faja de Oro Fault
1 MEXICAN RIDGES
Top Pliocene Top Miocene
2
Two Way Time (s)

3
4
5 B
6
Top Oligocene
7 10 km
Top Mesozoic
8
Annapurna Field
Thick overburden

Seafloor
Sediment thickness
Thin overburden

Eocene

0 10’s m 100’s m > 1km Cretaceous


Narrow fault/detachment zone Thick mobile shale C
Mobile shale thickness

SW NE
Extensional Domain
Maast.
TWT u/c
Transitional Domain
(s.)

2
Contractional Domain

Seabed
4
M. Aptian

Break-up
Unconformity
6
25 km
D
Fig. 10. Range of structural styles associated with shale detachments. A = Amazon Fan, (based on seismic section in Cobbold et al., 2004), B = Mexican Ridges (based on seismic
section in Ambrose et al. (2005), C = Krishna-Godavari Basin, India, (redrawn from Cairn Energy online seismic image), D = Orange Basin, De Vera et al. (2010), E = first order control
on structural style scheme for shale detachments based on the scheme proposed for salt tectonics by Tari et al. (2003).

folds, and longer duration of activity tend to be associated with detachments are plotted; their wide separation illustrates the
stratigraphically thicker, more long-lived deltaic systems. The considerable variation in basic parameters within the Type 1a
thickness and viscosity of the detachment also have a first-order DWFTBs.
control on structural style, (Simpson, 2009). Fold trains, or fold Offshore dipping detachments tend to be driven by steepening of
trains with break thrusts, and thrusted diapirs tend to be associated the slope due to hinterland uplift, and hence several episodes of
with thick mobile shale units (e.g. Mexican Ridges, Inner Fold Belt, deformation may affect a margin over tens of millions of years
Niger Delta Fig. 10B). Conversely imbricate thrusts, pop-up and (Hudec and Jackson, 2004; Rowan et al., 2004). Major deltas are a
triangle zones and associated folds tend to be found along discrete more continuous source of near-field stress due to differential
basal thrust or thin detachment zones (e.g. Corredor et al., 2005; loading and gravity sliding, particularly where sediment thickness
Cobbold et al., 2009; Fig. 10A, C, D). This application of the Tari et al. is great enough to cause a landward dip to the basal detachment.
(2003) scheme is further modified to capture one important aspect Loading by major deltas ultimately imposes offshore progradation
of shale detachments: that the detachment does not necessarily and offshore propagation of folds and thrusts. The two outstanding
only dip oceanward. In the case of the Niger Delta the DWFTB is examples of this are the Niger Delta and northern Gulf of Mexico.
underlain by a landward dipping detachment (Morley and Guerin, Conversely, the offshore dipping salt detachment systems along the
1996; Bilotti and Shaw, 2005). Fig. 13 shows the range of near field SW African margin show a propensity for late-state landward
stress-related shale detachments based on stratigraphic thickness, propagation of deformation (Hudec and Jackson, 2004; Jackson et
mobile shale thickness and detachment dip. Five examples of shale al., 2008; as described below).
54 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

30
Mean value of fold population wavelength (km)
4 values only given where 15 or more folds were measured
vertical bar = range of values

20
Fold wavelength (km)

16.7

15

12

10 10

7.9

5.6 5.7 5.4


4 4.2
3

0
Nile Delta (Messinian)

Makassar Straits
Mexican Ridges

Scotia Basin

Seram
Kwanza (Congo)

NW Borneo
Indo-Burma Ranges
Gabon (Astrid)

South Caspian Sea


Amazon Fan

Orange Basin

Cyprus
Angola

Makran
Niger Delta

Zagros
Perdido

Shale detachment Salt detachment Shale detachment

Type 1 Type 2
Fig. 11. Fold wavelengths for a range of DWFTB settings. The Zagros Mountains although not a DWFTB are illustrated for comparison with a classic orogenic fold and thrust belt.

4.1.2. Type 1a, DWFTBs associated with regional offshore-dipping extensional and compressional provinces “shale diapirs” or mud pipes
detachments along overpressured shales can be developed.
Examples: Bight Basin S. Australia (Totterdell and Krassay, 2003); A belt of regular, landward-dipping imbricate thrusts dominates the
Orange Basin S. Africa (de Vera et al., 2010); West Luconia Delta, W. DWFTBs of the offshore Bight Basin, S. Australia and the Orange Basin,
Borneo; Sandakan Basin, N. Borneo (Samsu et al., 2000); the NW flank South Africa. The width of the toe thrusts and folds in the Bight Basin is
of the Niger Delta, W. Africa, (Morley and Guerin, 1996). Krishna– 20–30 km, the thickness of section is up to 3 km, with ~3–4 km spacing
Godavari Basin (Annapurna Field, Cairn Energy); Mexican Ridges, Gulf between imbricate thrusts (Totterdell and Krassay, 2003). In the Orange
of Mexico (Ambrose et al., 2005), Rovuma basin N. Mozambique, Basin, toe folding and thrusting is spectacular, and occupies a zone up to
(http://www.hgs.org/en/cev/930/). 60 km wide in the transport direction, and lies within a section ~3 km
Despite displaying regional offshore dips, large tracts of passive thick (Fig. 10D). Spacing between the imbricate thrusts is 1–5 km. As is
margins do not exhibit well-developed gravity tectonics related commonly the case in thrust-associated deepwater folds, the back-limb
structural provinces with linked, large-scale extensional and contrac- dip of the anticline is gentler than the dip of the imbricate thrust (de
tional deformation provinces. However, extensive regions of nested, Vera et al., 2010). Hence the geometry does not conform to the classic
small-scale normal faults are commonly present and indicative of (no shear) fault bend fold or fault propagation fold model (e.g. Suppe,
widespread deformation related to mass movement, diagenetic 1983; Suppe and Medwedeff, 1990), but instead appears to be more
processes and dewatering (see for example Cartwright and Huuse, appropriately described by a shear fault-bend fold model (Suppe et al.,
2005). An offshore-dipping slope is generally insufficient to induce 2004), or a break thrust model (e.g. Eisenstadt and De Paor, 1987;
widespread, large-scale gravity sliding in shale-prone passive margin Morley, 1994, 2009b).
sequences; instead such systems occur where high sedimentation The up-dip extensional province in the Orange Basin is strongly
rates associated with deltas are present. Type 1a systems tend to segmented by lateral faults, parallel or sub-parallel to the transport
display relatively simple geometries with a system of predominantly direction (de Vera et al., 2010). This segmentation appears to
oceanward-dipping normal (regional) faults, that pass down dip into continue, although not necessarily along the same structures, down-
a zone of mostly landwards-dipping imbricate thrust faults and dip into the fold and thrust belt where it causes abrupt along-strike
oceanward-verging folds (Fig. 10). In the transition zone between the changes in the location of folds and thrusts (de Vera et al., 2010). Such
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 55

10 5

Stratigraphic thickness (km)


6

2 4
5 1 = Orange Basin
2 = Mexican Ridges
7 3 = Outer fold and thrust
3 belt Niger Delta
4 = Indo-Burma Ranges
5 = South Caspian Sea
1
= range of values 6 = NW Borneo
7 = Seram
= mean value 8 = Makassar Straits
0

0 5 10 15 20 25 30 km
Fold wavelength

Fig. 12. Wavelength vs. stratigraphic thickness of deformed series for eight shale detachment-associated DWFTBs.

segmentation is likely to exist in other, similar DWFTBs. The origin of 1983, Feng et al., 1994, Wawrzynice et al., 2003, Ambrose et al., 2005).
oblique and lateral faults is highly variable and for both salt and shale The ridges are predominantly offshore verging folds, with some
detachments and includes: inheritance from underlying structures, break-thrusts in their forelimbs. The folding style contrasts strongly
inheritance of lateral faults from up-slope, lateral changes in with the imbricate thrust style of many other Type 1ai DWFTBs
detachment zone thickness or material properties, lateral changes in (Fig. 10A, C, D). Probably this is due to a relatively thick zone of mobile
thickness or lithology in the detachment upper plate, the effects of shales forming the Oligocene detachment zone.
variable strain around an object in the lower plate, and conjugate It is expected that extension should approximately balance
strike-slip faults in type-1 fold-related fracture-type orientations, contraction in the toe area in linked systems. However, quantifying
(Stearns and Friedman, 1972), e.g. Lewis et al. (1988), Jackson et al. the contraction can be difficult since poorly lithified, water-rich
(2008), Morley (2009b), Clark and Cartwright (2009). sediments have been deformed and consequently simple line-length
The Mexican Ridges DWFTB in the western Gulf of Mexico is or area balancing of folds and thrusts will not necessarily account for
developed above an Oligocene shale detachment (Fig. 10B; Buffler, all the deformation. Structurally induced compaction and dewatering
may also play a significant role (e.g. Morgan and Karig, 1995; Henry et
al., 2003; de Vera et al., 2010; Butler and Paton, 2010). In the Orange
Basin, the difference between extension (24 km) and shortening
1 = Main Niger Delta (16 km) measured from large-scale structures visible on seismic lines,
2 = NW margin of Niger Delta suggests sub-seismic deformation processes may account for about
3 = Amazon Fan
24% shortening (de Vera et al., 2010).
4 = Mexican Ridges
5 = Orange Basin 1 Deformation along a shale detachment will tend to cause
3 dewatering and loss of the initial weak characteristics of the shale
10 layer particularly if disequilibrium compaction is the main mech-
Thickness of stratigraphy (km)

anism for generating overpressures. Consequently new deltaic


cycles on the margin, or new episodes of uplift may not cause
reactivation of the initial detachment, the margin may cease to be
prone to large-scale gravity sliding (as seen in the different
2 deformation styles of the Albian-Cenomanian White Pointer, and
5
5 Santonian–Masstrictian Hammerhead deltas of the Bight Basin,
Totterdell and Krassay, 2003). Alternatively several stacked detach-
4 ments of different age may be present where highly overpressured
shales are present, e.g. the Orange and Krishna–Godavari basins
(Fig. 10C and D).
0
10 -2° The Annapurna Field, with mean resources of 76 MMBOE, was
’s m
) discovered by Cairn Energy in the Krishna–Godavari Basin, offshore
De
tac
10
0 w ard eastern India. The basin contains one of the few fields found so far in
hm ’s m 0° an
oce
en
t th t d ip ( relatively small Type 1 shale detachment provinces. The field is
ick 10 en
ne 00 hm located in a broad anticline set up between the compressional toe
ss ’s m tac
10° De thrust and the extensional faults up dip (Fig. 10C). This is an atypical
Fig. 13. Triaxial scheme illustrating the wide range of basic parameters (thickness of
structure compared with the detachment, fault bend and fault
stratigraphy, detachment thickness, detachment dip direction) associated with Type 1a propagation folds found in the larger delta DWFTBs of the Niger
shale detachment DWFTBs. Delta and Gulf of Mexico.
56 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

4.1.3. Type 1a, DWFTBs associated with a widespread hinterland-dipping Formation, in particular due to variations in pore fluid pressure and
basal detachment zone (large deltas) sand content, are likely to be responsible for many of the changes in
The Niger Delta displays the classic, well-developed zones of a structural style, sequence of deformation and location of detachment
large delta province with an onshore to shelfal zone of growth fault- levels observed in the DWFTB (Corredor et al., 2005; Briggs et al.,
controlled depocentres, a well developed belt of “shale diapirs” 2006). Basement topography influences the DWFTB structure in
around the outer shelf-slope area, and a fold and thrust belt in the places too, for example in the location of lateral ramps, and the
slope area (e.g. Evamy et al., 1978; Doust and Omatsola, 1989; Morley presence of basement highs in front of well-developed imbricate
and Guerin, 1996; Haack et al., 2000; Ajakaiye and Bally, 2002; structures (e.g. Morley and Guerin, 1996; Morley 2003a; Corredor et
Fig. 14). The Cenozoic stratigraphy of the delta is divided into three al., 2005).
main formations (e.g. Evamy et al., 1978), the Benin Formation Corredor et al. (2005) divided the DWFTB into an inner fold and
(alluvial), Agbada Formation (deltaic-inner shelf mixed sands and thrust belt characterized by oceanward verging, commonly imbricat-
shales), and the Akata Formation (outer shelf, slope, and bathyl ed, thrusts and folds, including detachment folds, a transition zone
deepwater shales with deepwater sands). For the deepwater area the which is largely undeformed except for large, broad folds, and an
Akata Formation contains source, seal and reservoir rocks. outer fold and thrust belt with basinward- and hinterland-verging,
The delta province extends about 300 km from the onshore conjugate thrust faults and associated folds (Fig. 15). Many of the
extensional province to the thrust front of the DWFTB, and covers a structures are active and their geometry is reflected in the
distance of about 500 km along strike (Fig. 14). The DWFTB forms two bathymetry. The outer fold and thrust belt has been the main focus
lobes, each about 200 km along strike, with a central area that is of deepwater exploration, in particular the southern lobe. For example
largely unaffected by thrusting and folding that coincides with the the Agbami Field (1998; Fig. 14) is located in the southern lobe, and is
Charcot Fracture Zone (Fig. 14). The DWFTB extends about 80–100 km the largest of the deepwater discoveries in Nigeria, with estimated
in the transport direction, and has undergone maximum shortening of recoverable volumes of 900 MMBBO, and production rates of
about 17–25 km (Corredor et al., 2005). Some folds display pop-up 250,000 BBO/day (Maksoud, 2008).
and triangle-zone geometries, or even a series of back-thrusts (e.g. The low-angle of the basal detachment and fore-and back-thrust
Bilotti and Shaw, 2005), but, in general, most folds and thrusts verge fold style in the outer fold and thrust belt (Fig. 15) implies a low
offshore. The Niger DWFTB forms a low critical-taper wedge with a resistance to basal slip and high pore fluid pressures for the sharp
1°–2° surface slope (α) and 1°–1.5° basal detachment dip (β) with an detachment zone that lies within the Akata Formation (Cobbold et al.,
overall wedge taper of (α + β) 2.5° ± 0.4°, which is explained by high 2009). Relatively slow seismic velocities within the Akata Formation
pore fluid pressure ratios (λ = ~0.9) along the basal detachment support this inference of high pore fluid pressures, and a pressure
(Bilotti and Shaw, 2005). Typical for a critical wedge model, the ramp near the top of the formation observed in wells (Morgan, 2003;
sequence of thrusting tends to young offshore, but there are instances Cobbold et al., 2009). Frost (1996) and Cobbold et al. (2009) have
of out-of-sequence thrusting, synchronous deformation or break-back argued for that hydrocarbon generation triggered by burial within the
sequences on a number of thrusts and folds (Morley, 2003a; Corredor thrust belt is responsible for the large-volume overpressures. The
et al. 2005). Spatial and temporal changes in the strength of the Akata cycle of burial by thrusting and sedimentation that lead to

0 50km Folds

Edge of thrust
Thrust fault
and fold belt
Normal fault
Transition zone Erha Field
Diapir ridge axis

5° Fig. 15 5°
Outer thrust
and fold belt
Bonga Field

4° Inner fold and thrust


belt, including thrusted

Fig. 17
diapris
ge
Diapir province
ot Rid Diapir belt at
arc
Ch Fig. 16 thrust front
Zone dominated by Agbami Field
regional growth faults
3° Outer thrust 3°
Zone dominated by counter and fold belt
regional growth faults Oblique thrust
Akpo Field at basement fault
Transition zone trend
4° 5° 6° 7° 8°

Fig. 14. Regional map of the main structural provinces of the Niger Delta (modified from Morley and Guerin, 1996 and Cobbold et al. (2009).
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 57

SSW Ocean floor NNE

Agbada Fm.

Akata Fm.
1 s. TWT 2 km

Fig. 15. Outer fold and thrust belt of the Niger Delta, drawn from 2D seismic line in Cobbold et al. (2009). See Fig. 14 for location.

overpressure generation probably caused a feedback mechanism that pressured shales that thin oceanward. Consequently, earlier up-dip
aided oceanward propagation of the thrust front (Cobbold et al., 2009). extension was mostly accommodated by movement of thick mobile
Evidence that a significant portion of the up-dip extension is shales to form the “shale diapir” belt (Fig. 16B, C; see discussion in
accommodated by shortening landwards of the present DWFTB Section 3.4, and Ings and Beaumont, 2010).
includes: 1) the relatively young age of most of the folds (Late Once the delta began to prograde over the “diapir” belt, the DWFTB
Miocene–Holocene) compared with the Eocene–Holocene age of the started to develop oceanward of the “diapirs” with a relatively thin
delta. 2) Isolated examples on the shelf of individual growth fault basal detachment zone in most places (Morley, 2003a). In Fig. 16D the
depocentres that have self contained extensional–contractional latest stage sedimentary sequence (mid Pliocene–Holocene) displays
systems with growth faults passing down-dip into either “shale a broadly synformal geometry in the domain of the “diapir” province,
diapirs” or an individual toe thrust (Fig. 16), and 3) The extensive and is not strongly fault controlled. This geometry suggests a broad
“diapir” belt that lies between the DWFTB and the extensional subsidence of the basal mobile shale layer to accommodate the
province has taken up some horizontal extensional displacement by sedimentation, possibly largely achieved by fluid loss. The geometry
vertical motions and fluid expulsion (Morley, 2003a). Progradation of implies broad differential loading, rather than extensional faulting
the delta over older, more landwards fold and thrust belts possibly may have driven the later stage of deformation in the outer fold and
caused them to be lost into the deeper, poorly imaged section. thrust belt in the southern lobe.
However, it is also likely that the structural style has changed with Improvements in seismic data quality commonly show that chaotic
time as the sand-prone delta prograded over thick mobile over- regions interpreted as diapirs on older vintages of seismic data shrink in

Outer fold and Inner fold and Late stage synformal geometry
S thrust belt thrust belt to Mid Pliocene-Holocene section N

Mid Pliocene-Holocene Mobile shales Mobile shales, or partially 10 km


mobile shales, or just
Near top Miocene-Mid Pliocene
poorly imaged stratified shales?
Palaeogene-Near top Miocene

Seafloor

A
Mobile shale limit of strong overpressured,
mobile shale

Seafloor
Migration of
overpressured fluids

B
Upwards moving front of overpressure
Late stage synformal geometry
to Mid Pliocene-Holocene section
Seafloor

Fig. 16. A) Cross-section based on 2D seismic reflection data (Ajakaiye and Bally, 2002) across the southern DWFTB lobe of the Niger Delta (see Fig. 14 for location). B) Conceptual
evolutionary diagram of southern Niger DWFTB: A) Early stage of development with movements on growth faults passing down-dip into shale ridges, B) seaward propagation of
growth faulting and development of second, thrusted shale ridge. C) Cessation of growth faulting, collapse and movement on mobile shale due to differential loading in a saucer-
shaped basin (only minor fault control). During this differential loading stage the outer fold and thrust belt developed.
58 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

0 shale with a diapiric geometry maybe involved out to the thrust front in
W E
places. The eastern termination of the southern lobe is a narrow feature,
with a very different style from the relatively widely spaced imbricates
of the central southern lobe (e.g. Corredor et al., 2005; Cobbold et al.,
2009). Imbricated slivers are stacked much more vertically and a narrow
zone of mobile shale is possibly involved in the imbricate stack (Fig. 17).
In a different deepwater setting Barber et al. (1986) have shown

TWT (s.)
from outcrop that large chaotic masses of squeezed diapirs form in
overpressured DWFTBs, and may be an important mechanism of
mélange formation. It is reasonable to suspect that such processes
5 could also operate within the DWFTB of the Niger Delta. At the very
least there is a cycle where very large volumes of fluid and mud are
pumped from the deeper parts of the delta to the surface along
vertical pipes to form mud volcano fields, and then the mud becomes
re-deposited within the delta sediments and buried (mud pipe-
10 km volcano systems may extrude volumes of material in the order of 1–
8 11 km2 over their lifetime, Guliev, 1992; Graue, 2000). This cycle,
Upper syn-kinematic sequence Middle syn-kinematic sequence coupled with significant shortening by lateral, compression-related
compaction cast doubts on the utility of regional balanced cross-
Upper pre-kinematic sequence Lower syn-kinematic sequence
section solutions for entire deltaic provinces that treat deformation
Mobile shales (Akata Formation) Lower pre-kinematic sequence only in terms of linked fault systems, narrow detachment faults,
simple fault-related folds and assume constant area or line length (e.g.
Fig. 17. Stacked imbricates and mobile shale eastern Niger Delta, based on Ajakaiye and
Butler and Paton, 2010).
Bally, 2002, their Fig. 1 profile D3; see Fig. 14 for location).
The Niger Delta is larger, and contains more structural variety than
the shale detachment systems described above and summarized in
size, or are replaced by layered sequences on newer data (e.g. Van Fig. 10. In particular, the presence of large “diapirs” and an external
Rensbergen et al., 1999; Van Rensbergen and Morley, 2003). The DWFTB thrust belt developed above a hinterland-dipping basal detachment
of the Niger Delta contains folds and imbricates that in places clearly (e.g. Evamy et al., 1978; Morley and Guerin, 1996). This sets the Niger
only affect strongly reflective sequences, but in other places are weakly Delta apart from other near-field stress driven Cenozoic deltas. One
reflective or chaotic (Ajakaiye and Bally, 2002). Some recent publica- other exception may be the McKenzie Delta, but its structure is not
tions (e.g. Corredor et al., 2005; Kostenko et al., 2008) have emphasized well described in the literature and so the Niger Delta is discussed
the regular thrust belt geometries that clearly occur in many parts of the here as a lone example.
DWFTB. However, in parts of the delta, particularly around the outer
diapir province and the eastern margin of the DWFTB, there is a different 4.1.4. Type 1b, DWFTBs associated with an oceanward-dipping salt
structural style, where mobile, overpressured shale diapirs seem to form detachment zone
the core to imbricated structures (Fig. 16; Morley, 2003a). Ajakaiye and We estimate that thick salt units are found along about 13,000 km
Bally (2002) show the potential for mobile shale involvement within the length of the world's passive margins. The most important areas are
fold and thrust belt, in particular their Fig. 1 profile D3 suggests mobile the Atlantic margins of South America (Campos, Espirito Santo,

Outer Kwanza Basin (passive margin) Inner Kwanza Basin (interior basin)
Outer Salt Basin Flamingo Platform Inner Salt Basin Eastern Rim
Abyssal Angola Salt Monocline domain
Thickened salt palteau Diapir domain Raft Domain Mock Turtle Domain Updip Wedge
Plain Nappe
Atlantic Hinge Zone Coast line
A WSW ENE
0

2
Depth (km)

4 Congo Craton

6
Distal Ramp 50 km
Aptian salt
8 Probable Gabon-Angola horst and graben system
Truncated shelf
B SE Thin-skinned extension NW C S Contractional domain Extensional domain N
Thin-skinned contraction 0
Inverted extensional
10 km
Two-way time (s.)

3.0 Complex fold-belt 1 domain


4.0 Astrid thrust belt
5.0 2
6.0 Albian- Distal wall
7.0 3 Sea floor 20 km
8.0 Cenomanian
9.0 Upper and Middle Jurassic 4
TWT (s) salt-cored anticlines, forced folds above basement blocks 5
6
Aptian salt
7
Allochthonous salt fringe Basement ramp

Fig. 18. Cross-sections across Type 1b (salt detachment) margins. A) Regional cross-section across the Kwanza Basin, Angola, redrawn from Hudec and Jackson (2004). B) Cross-
section across the Sable sub-basin, offshore Nova Scotia based on seismic reflection data, (Deptuck et al., 2009). C) Regional cross-section from the inner continental shelf to the
deepwater Astrid thrust belt, Gabon, redrawn from Jackson et al. (2008).
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 59

4 km of salt, capped by a thin veneer of sediment, was present in the


western part of the basin at the time of initial deformation. Passing
eastwards, up to 3–4 km thickness of sediment overlay thinner, more
isolated salt bodies. The isolated character was due either to salt
movement, or salt depositional thickness not exceeding 50–150 m.
The Kwanza Basin is divided into two sub-basins, the Inner Kwanza
Basin (predominantly located onshore at present-day) and the Outer
Kwanza Basin (located on the shelf to deepwater). A broad high
separated the Inner Kwanza Basin from the Outer Kwanza basin for
most of the Cretaceous–Cenozoic, so there was no linked detachment
system between the two basins for most of their history. This
description focuses on the Outer Kwanza Basin, which had a DWFTB
linked up-dip to an extensional system throughout its history.
The presence of the thick salt caused the structural styles in the
DWFTB to be much more varied than the type 1a DWFTBs, and include
a large-scale salt nappe, thickened salt plateau, and folded and
thrusted (squeezed) diapirs (Figs. 18 and 19). Above the salt nappe,
are relatively short wavelength (2–3 km), broad, open folds affected
by both normal faults and thrusts, or with no faults at all. There is no
strong vergence to the structures.
According to Hudec and Jackson (2004), three phases of regional
uplift are each associated with up slope extension and down slope
contraction, but the location and structural style of the deepwater
zone of contraction differs for each phase. The Albian phase occurred
over a wide area and involved short-wavelength buckle folds (up to
~5 km) and compressional diapirs. Most shortening occurred up-dip
of the basinward salt pinch-out. The Late Cretaceous–Middle Miocene
contraction was focused on the basinward limit of the salt by advance
of the Angola salt nappe. There appears to be little contraction within
synclinal mini-basins created during Albian buckling. About 20 km
basinward translation affected the Outer Kwanza Basin during the
Oligocene–Holocene. Miocene deformation is marked by acceleration
in nappe translation, and the contractional zone broadened to a 150-
km wide zone from the leading edge of the allochthonous salt to the
base of the continental slope. Hudec and Jackson (2004) estimate that
relatively minor uplift and tilting of the continental shelf resulted in
an order of magnitude increase in translation rate (from 300 to
3200 m/my).
Increased sedimentation rates from 8 to N130 m/my in the
Fig. 19. Conceptual cross-sections showing the range of salt involving compressional abyssal-plain region occurred during the Late Miocene. One effect
structures evident in the Astrid, DWFTB, Gabon, (redrawn from Jackson et al., 2008). was to lower the slope on the continental rise and thus decrease the
A) Thrusted low-amplitude salt anticline, B) small buried diapir with an overhanging potential energy driving the Angola salt nappe. A second effect was
bulb, and C) precursor diapir with an allochthonous sheet and pinched-off feeder. The
aggradation of sediment oceanward of the distal salt margin. The roof
growth stages are described in the right margin.
of the nappe became effectively buttressed, and its advance hindered.
Consequently oceanward-propagating deformation was limited and
Sergipe and Santos basins), West Africa and Canada (Scotia Basin), the instead there was a landward shift in shortening (Hudec and Jackson,
Gulf of Mexico, and the Red Sea. Unlike the Type 1a margins, which 2004). The landward propagation caused older salt structures to be
are associated with a shale detachment, gravity-driven deformation reactivated and laterally shortened. Some reactivated structures
above a salt detachment can occur outside of areas with large deltaic display a strong influence of sediment package geometry developed
input onto the margin. Commonly, uplift of the continental margin under extensional or diapiric conditions.
triggers salt movement. For example, the Angolan margin underwent
three phases of uplift (early Albian, Campanian and Miocene) that 4.1.4.2. Lower Congo Basin, Astrid thrust belt. In many parts of the
triggered reactivation of the same Aptian salt layer (e.g. Hudec and Congo passive margin Basin relatively thick salt has led to the
Jackson, 2004). Structures associated with mobile salt can be very development of allochthonous salt canopies and structures akin to the
complex, and vary depending upon the thickness of the salt layer, Kwanza Basin (Marton et al., 2000; Tari et al., 2003). However, the
geometry of pre-existing structures, and uplift history and sediment Astrid Thrust Belt of the Lower Congo shelf displays deformation
loading history of the margin (Fig. 18). associated with a much thinner salt layer, and relatively thicker
Five basins are described below to illustrate the primary variations sedimentary section (Congo Fan) than the Kwanza Basin described
in margin setting, salt thickness, salt timing (syn-rift or post-rift above (Fig. 18C; Basin; Jackson et al., 2008).
deposition), and triggering mechanism (thermal subsidence, hinter- The following account is based on Jackson et al. (2008). Gravity
land uplift, deltaic loading). driven shortening during the Aptian caused the development of
gentle, west-trending salt-cored anticlines. The main thrust belt
4.1.4.1. Kwanza Basin, Angola. The Kwanza Basin is an outstanding formed during the Late Cretaceous, when a landward propagating
example of a long-lived thick salt mass on a passive margin. The thrust belt developed in response to epirogenic uplift. The Astrid
following description is based on Hudec and Jackson (2004). The thrust belt is associated with a detachment that dips oceanward
thickness of the deformed section is highly variable (Fig. 18A). Up to between 1° and 3° and extends over 130 km in the transport direction.
60 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

The importance of pre-existing structures is emphasized in this thrust section that exceeds 10 km in places and both salt and shale
belt because the salt layer was relatively thin at the time of detachments are present. Underlying the system is the oceanward-
contraction, so pre-existing salt diapirs and gentle anticlines influ- dipping the Louann Salt basal detachment (Fig. 20).
enced the contractional structures (Fig. 19), resulting in a regular Depositional loading during the Cenozoic triggered the main
spacing and strike to the thrusts, and curvature of thrust tips to link episodes of gravity deformation in the Gulf. Changing patterns of
with pre-existing diapirs. Late expulsion of salt from buried anticlines deposition in the Gulf have been identified in a number of studies
caused folding of overlying thrusts in some places. There is a strong as reviewed by Galloway et al. (2000), and show that the southern
seaward vergence due to the commencement of thrusting after Rocky Mountains were the main sediment source area during the
welding had occurred along the basal salt. Like the Kwanza Basin, the earlyPalaeogene, and the Houston embayment was the main
landward edge of the thrust belt propagated landwards, probably due depocentre (Fig. 21). Uplift in northern Mexico of the Sierra
to buttressing down dip by the overburden. During thrusting, early Madre and Trans-Pecos in Texas increased sedimentation in the Rio
passive diapirs were squeezed and extruded to form small allochtho- Grande embayment slightly later (upper Wilcox). The major
nous sheets (Fig. 19C). The propagation of shortening up-dip towards depocentre shifted between the Houston and Rio Grande embay-
the shelf into previously extensional domains resulted in inversion of ments until the Middle Miocene, after which time drainage shifted
older normal faults. eastwards to the Mississippi Embayment as a result of uplift of the
Western Interior.
4.1.4.3. Scotia Margin. The Scotia Margin is noteworthy in its different The deepwater area of the Northern Gulf of Mexico is critical to
behavior compared with the West African examples described above. sustained large-scale hydrocarbon production from the region.
First, the salt is of Triassic age, deposited during the syn-rift stage DWFTB plays are only one of three main exploration plays in the
(Wade and MacLean, 1990), hence there is a strong influence of the deepwater province, the other two being the Flex Trend and Mini-
rift tilted fault block geometry on the location of folds, diapirs and Basin trends (Cossey, 2004). Deepwater production as a percentage of
faults (Fig. 18B). Secondly, deformation was initially driven by delta total Gulf of Mexico production has risen from ~ 6% Oil (21 MMSTBO
loading during the Middle and Upper Jurassic syn-rift stage (Ziegler, and 0.8% gas in 1985 to 73% oil (308 MMSTBO) and 43% gas (~ 1 TCFG)
1989). Most of the extension in some places has been accommodated in 2008. According to Nixon et al. (2009), over 141 producing projects
by extrusion of a salt nappe at the sea floor, with little development of existed at the end of 2008 in the deepwater Gulf of Mexico.
a contractional zone (Ings and Shimeld, 2006). However, an up dip There are three main DWFTBs in the Northern Gulf of Mexico, the
extensional system is balanced by a broad zone of down-slope Perdido, Mississippi Fan and Port Isabel fold and thrust belts (Fig. 21).
contraction in the Sable sub-basin, in a region between a major salt Reviews of the deepwater exploration in these trends are provided by
canopy system to the east, and isolated salt diapirs to the west Meyer et al. (2005, 2007). A cluster of deepwater hydrocarbon
(Fig. 18B, Deptuck et al., 2009). development projects (BAHA, 2002; Trident, 2001; Great White, 2002;
Silvertip, Tobago) exist in the Perdido Fold Belt, but more numerous
4.1.4.4. Northern Gulf of Mexico. The Gulf of Mexico is a very large, development projects are associated with the Mississippi Fan Fold Belt.
complex and very well documented near-field stress-driven system. Major discoveries in the Mississippi Fan Fold Belt include Neptune
In this kind of review it is impossible to do justice to the depth of (1995), Mad Dog (1999), Tahiti (2002), Knotty Head (2005), Genghis
information and understanding that exists for the Gulf, instead some Khan (2005) and Big Foot (2005). Thunder Horse (1999) is the largest of
highly generalized comments must suffice. the DWFTB fields, with an estimated mean size of 1 billion barrels of oil
The Gulf is an unusually long-lived gravity driven system, in terms equivalent (BBOE) (Cossey, 2004). In March 2009, the field was ramping
of size, longevity, mixture of salt and shale mobile units and resulting up to produce 300,000 BOE per day from only seven wells. It is the
structural complexity, it can be regarded as the most complex end- largest producing deepwater asset in the world (http://www.bp.com/
member of the near-field stress systems. The uniqueness of the Gulf genericarticle.do?categoryId=9004519&contentId=7009088, http://
lies in its tectonic setting: the basin opened by continental extension www.businessweek.com/magazine/content/09_37/b4146000578301.
from ~ 160 to 150 Ma, followed by a short period of seafloor spreading htm). Conversely, the Port Isabel Fold Belt has no discoveries (Cossey,
from 150 to 140 Ma and then became inactive (see review in Bird et 2004).
al., 2005). Seawater influxes during the syn-rift stage permitted The northwestern Gulf of Mexico contains the deepwater Perdido
deposition of the massive Middle Jurassic Louann Salt (e.g. Peel et al., fold and thrust belt. This belt is, however, just the external portion of a
1995). The Northern Gulf of Mexico contains a great thickness of more extensive belt of contraction that includes the Port Isabel fold

Fig. 20. Cross-section across the northwestern Gulf of Mexico, redrawn from Peel et al. (1995) and Trudgill et al. (1999). Gravity-driven deformation occurs along a complex system
of salt and shale detachments.
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 61

95°W 90°W

e New Orleans
om
lt D Little Tertiary 30°N
Sa Mi
ni extension or
Ba compression
Houston s in
Pro
ce vin

s
in

m
ce
USA ov

te
s Pr Salt Dome-Mini Basin

ys
tem t

tS
Sy
s en Province

ul
m
ult ch
Fa
a ta
xF ne
i lco Eo
ce De
per e
W
Up en Plio-Pleistocene Detachment Province
ioc
t
en

M
o-
hm

ig
Ol Tabular Salt Mini Basin Province
tac
De

Port Isabel
Mississippi Fan
urg

Fold Belt
ksb

Fold Belt
Vic

Local collision of salt-


withdrawal basins US Waters

Sigsbe nt Mexican Waters


e Escarpme
Mexican Perdido 25°N
Ridges Fold Belt
Fold Belt -3000 m

-200 m
-3000 m
MEXICO

-200 m

0 300 km

Fig. 21. Map of the Gulf of Mexico showing main structural units and water depths, based on maps in Trudgill et al. (1999) and Ambrose et al. (2005).

belt and the salt canopy area (Fig. 20; Peel et al., 1995; Trudgill et al., Late Jurassic–Cretaceous in response to early contractional deforma-
1999). Up-dip extension of 60 + km suggests the down-dip fold and tion, and later helped control the Cenozoic age folding (Grando and
thrust belts, and the salt canopy must accommodate a similar amount McClay, 2004). The fold belt is partially overlain by the Sigsbee Salt
of contraction, of which the Perdido fold belt only accommodates 5– Nappe, whose emplacement is responsible for the withdrawal of salt
10 km (Peel et al., 1995; Trudgill et al., 1999). The Perdido fold belt along the basal detachment of the Mississippi Fan Fold Belt. This
mainly underwent deformation during the lower Oligocene, but situation contrasts with the thick salt detachment of the Perdido Fold
minor deformation continued until the Middle Miocene (Trudgill et Belt (Fig. 22; Trudgill et al., 1999).
al., 1999). In map view, the folds range from linear, to more dome
shaped, fold length appears to be limited in places by underlying NW– 4.1.4.5. Nile Delta. The eastern Mediterranean in the Levant region of
SE trending syn-rift basement structures (Trudgill et al., 1999). The the African Plate underwent rifting during the early Mesozoic
fold belt underwent a late oceanwards tilting as a result of a wedge of (Gardosh and Druckman, 2006). Extension probably ceased prior to
salt being injected into, and inflating, the basal detachment (Fig. 22; the creation of oceanic crust, but was accompanied by extensive
Trudgill et al., 1999). A complex system of extensional detachments volcanism. Subsequently, during the late Cretaceous to Miocene
exists up-dip, with detachments forming at different levels partly in several phases of inversion affected the older normal faults giving rise
response to large salt sheets migrating up through different to the Syrian Arc fold belt. The Messinian evaporites that form the
stratigraphic levels over time (e.g. Diegel et al., 1995). upper detachment in the deepwater Niger Delta arose as a result of
The switch in depositional systems with time resulted in isolation of the Mediterranean basin during plate convergence
abandonment of the Perdido fold belt, and formation of the between African and Europe (see review by Jolivet et al., 2006). At
Mississippi Fan fold belt during the Miocene–Pliocene (Wu et al., present the African Plate is being subducted northwards along the
1990; Weimer and Buffler, 1992). The two fold belts are separated by Cyprus Arc or undergoing sinistral motion along that trend (Cavazza
a wide region without demonstrable contractional deformation et al., 2004; Robertson and Mountrakis, 2006).
(Diegel et al., 1995). The Mississippi Fan Fold Belt is characterized The Nile Delta is a large delta that was activated during the Late
by broad, upright folds to asymmetric, offshore and landward verging Miocene (Tortonian) in conjunction with uplift of the Red Sea-Gulf of
folds, cut by thrusts with wavelengths of about 8–12 km. They are Suez rift shoulders (Guiraud et al., 2001) and is associated with
salt-cored detachment anticlines, with welds at the base of the deformation along two main offshore-dipping detachments. The
synclines (Rowan et al., 2004; Grando and McClay, 2004; Fig. 23). deeper detachment occurs in shales, while the upper one is in
Precursor, small wavelength salt pillows formed in places during the Messinian evaporites (Aal et al., 2000; Dolson et al., 2000; Tingay et
62 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

Fig. 22. Cross-section illustrating the development of the Perdido fold belt, Gulf of Mexico, by regional gravity spreading linked to up-dip extension. A) Paleogene initial loading and
extension caused primary extrusion of salt and the development of a regional allochthonous salt canopy. B) Closure of the feeder salt diapirs transferred stress to the toe of the
system, causing development of the Perdido fold belt. C) Further salt movement resulted in inflation of a salt wedge along the Perdido fold belt detachment. Complex deformation in
the Paleogene salt canopy is not represented. Redrawn from Trudgill et al. (1999).

al., 2010,in press). This stacking sequence is due to the location of the 4.2. Type 2, continental convergent tectonic setting
delta on a passive margin situated on a northward subducting lower
plate. This convergent setting resulted in the deposition of evaporites Continental convergent zones contain DWFTBs affected by both
during a late stage of passive margin development as the basin began near- and far-field stress systems (type 2a), and type 2b DWFTBs, which
to close, as opposed to the classic passive margin sequences seen on are entirely or predominantly affected by far-field stress systems. The
the Atlantic margins, where halites were deposited during basin island of Borneo is a remarkable area for deepwater exploration,
opening. Up-dip extensional faults and down-dip squeezed diapirs encompassing five different deltaic provinces with deepwater potential
and folds are associated with the Messinian evaporite detachment or proven fields that formed in a variety of Type 2 DWFTB settings. These
(Loncke et al., 2006). These authors show the whole systems is about provinces are the West Luconia, Sandakan, Tarakan and Mahakam
200–250 km long in the transport direction, with the compressional deltas, and the NW Borneo Trough (Brunei–Sabah) (Fig. 4). In addition,
zone being about 50–100 km wide. The structures display a wide the west Sulawesi Fold Belt lies opposite the Mahakam Delta on the west
range of orientations particularly in the extensional province, which Sulawesi margin in the Makassar Straits. Of these provinces the NW
reflect the interplay between pre-Messinian topography and the Borneo Trough, Mahakam and Tarakan deltas all have proven
distribution of Messinian evaporites, varying slope gradients at the deepwater hydrocarbon fields. The proliferation of deepwater prospects
base of the salt layer, and major features such as the Eratosthenes in the area is due to the uplift of Borneo during the Miocene, which
Seamount in the northeastern part of the deepwater delta (Loncke et resulted in rapid erosion of the uplifting island and rapid deposition
al. (2006). along the fringes (e.g. Hall and Nichols, 2002).

NW SE
Sea floor

4.0
Late Miocene Unconformity
5.0
TWT (s.)

6.0
Salt Canopy

7.0

8.0

Louann Salt 5 km Lower Cenozoic Wilcox Top Cretaceous

Fig. 23. Section across the Mississippi Fan Fold Belt, based on seismic line in Meyer et al. (2005).
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 63

Complex, early-stage continent–arc collision is ongoing in the In this section, the DWFTBs associated with Borneo and Sulawesi
Banda Arc area, east of Borneo, where the Australian Plate is colliding are discussed; these regions may, however, differ from classic collision
with the arc system at the eastern end of the Sumatra–Java subduction zones. These DWFTBs lie within a region of hot, weak continental
zone (Fig. 24). This area has reached the stage where the deepwater crust accreted during the Late Palaeozoic–Cenozoic that is unlikely to
thrust and fold belts are starting to be uplifted and eroded, and the deform in the typical style of cold, older continental crust (Hall and
deepwater troughs are being reduced in extent and depth due to Morley, 2004). Hall (in press) observes that, for Borneo and Sulawesi,
continued convergence and crustal thickening. Smaller, DWFTBs exist there was initial collision and shortening, but that later uplift and
in the region too, and appear to be associated with strike-slip erosion of the island interior was not accompanied by sufficient
deformation arising from collision (e.g. the Banggai-Sulu and shortening to drive the observed uplift. Consequently Hall (in press)
Cenderwasih DWFTBs, Ferdian et al., 2010; Sapiie et al., 2010; links sediment accumulation in basins fringing the island to uplift of
Fig. 3). The small size of the Cenderwasih DWFTB is particularly its interior in response to crustal-scale gravity-driven processes. There
remarkable, it has a (NE–SW) strike-length of only about 100 km, a appears to be a feedback mechanism between sediment loading in the
width of 30–40 km and lies in water depths up to 2000 m. The DWFTB fringing basins, and uplift of the island interior possibly involving the
is developed in a zone of intracontinental convergence where a flow of hot, ductile crust from beneath the basins towards the
Neogene foredeep basin overlying an older Permo-Triassic syn-rift uplifting island interior, which acts as the sediment source areas.
and Mesozoic post-rift sequence was folded and thrusted during the Deformation in the DWFTBs can be caused both by near-field gravity
Late Miocene–Holocene (Sapiie et al., 2010). processes limited to the sedimentary basin triggered by sediment

Pacific Ocean
126
Haimahera


17
Molucca
Sea
Bird’s Head
Banggai-Sula Misool Irian Jaya
Platform 32 77 Cenderawasih
Sulaweisi Ser Bay
am
Tro
ug h
Seram 100

4° Ambon
rc
A
da

Aru
an

st Tanimba
hru
B

Islands
gh

T
tar
u

We
ro
rT


ba
m
ni

Timor Arafura
Ta

Sea

u gh
Tro
or
Tim
Timor
Sea

12° 77
Australia
122°E 126°E 130°E 134°E

Thrust front of deepwater


GPS defined plate convergence
thrust wedge (early collisional thrust Arc volcanics
rates in cm/yr
wedge or accretionary prism)

Water depths between Water depths greater -2000 m sea -4000 m sea
sea level and -500 m than -4000 m floor contour floor contour

Fig. 24. Regional tectonic map of eastern Indonesia, Banda Arc area (partly based on Hall and Wilson, 2000; Sapin et al., 2009 and Ferdian et al., 2010).
64 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

loading of an overpressured mobile shale (Setap Shale), and by deeper trench-like morphology, the present day setting is not one of active
crustal and lithospheric far-field stresses. This crustal-scale coupled subduction. The presence of a train of deepwater folds along the NW
sediment loading-uplift model is similar in principal to the model for dipping slope was established from 2D data in the 1980's (e.g. James,
post-rift subsidence proposed by Morley and Westaway (2006) for 1984; Hinz et al., 1989). The fold belt is about 300 km long and 50–
another hot, weak part of the Sundaland crust in the Gulf of Thailand. 80 km wide and runs from the Baram Delta Province of the Brunei and
The model advanced by Hall (in press) needs further evaluation but northern Sarawak offshore area in the south, along the Sabah margin
represents an interesting possibility. Because this model requires towards Palawan (Fig. 25). The NW Borneo shelf lies to the SE of the
collisional process to initiate the orogenic zones and subsequent flow DWFTB and can be divided into two provinces: 1) the Baram Delta
of ductile crust, these processes can be regarded as deep seated in Province in the SW (locations 2 and 3, Fig. 25), and 2) the Sabah Basin
origin. Consequently these regions remain grouped with convergent in the NE (locations 4 and 5, Fig. 25). The two basins are transitional to
zones, and far-field stresses. Nevertheless, the model contains distinct each other and share the same characteristics, but to differing degrees.
differences with classic orogenic belts, particularly as the later stage The Baram Delta Province is dominated by a thick shelfal deposits
development does not require plate boundary forces to drive the affected by classic gravity-driven structures (overpressured shales,
deformation. growth faults, mud pipes). Conversely the shelfal area of the Sabah
The quality of clastic reservoir rock is strongly related to the tectonic Basin, while preserving some thick depocentres, growth faults and
setting and is generally a much higher risk for active margins and overpressured shales, is widely deformed by thick-skinned thrusts
collision zones than it is for passive margins (e.g. Dickinson and Suczek, and strike-slip faults that have caused multiple unconformities,
1979). For this reason the regional setting for sedimentation is discussed inversion structures, folding, uplift and erosion (e.g. Levell, 1987;
for Type 2 systems, while it was ignored for Type 1 systems. Tan and Lamy, 1990; Madon et al., 1999).
During the early Cenozoic, the Proto-South China Sea oceanic crust
4.2.1. Type 2a convergent zones was subducted to the SSE below NW Borneo, and the Cenozoic
Crocker Formation developed in an accretionary prism setting (e.g.
4.2.1.1. Baram Delta Province and Sabah Margin, NW Borneo. The Sabah James, 1984; Levell, 1987; Hutchison, 1996; Sandal, 1996; Hall, 2002).
margin, NW Borneo is marked by a NW dipping slope that passes into Jamming of the subduction zone by the entry of thinned Eurasian
a deepwater area more than 3000 m deep that marks the NW Borneo continental crust (Dangerous Grounds) during the early Miocene
trough. There is little earthquake or volcanic activity, so despite the terminated subduction and caused uplift and erosion of the area that

113 ° 114° 115° 116°

1 = Deepwater fold and thrust belt


gh

South China
ou

Seas 4
0


Tr

20

2 = Extensional growth faults in


5
eo

Baram Delta Province


rn

Offshore 3 = Inverted zone of Baram


Bo

Brunei Delta Province


W

4 = Mobile shale deformation province,


m

Darussalam
N
m

00
00

Sabah Basin
20
25

5 = Zone of strong inversion,


1
+ ++
m

Sabah Basin
00
10


2 σ1 orientations
from borehole breakouts

A quality
3 B quality
C quality
Mt Kinabalu D quality
5° granite
Sabah Anticline Syncline
B un
Brunei
n Normal Thrust
fault fault
Bathymetry

Zone of uplift
4° Sarawak 0 40 km including Kinabalu
Sarawak culmination
Meligan Sandstone
Paleogene Rajang Undifferentiated Paleogene Miri Zone,
Formation (Oligocene-
Group flysch (on oceanic crust?) (on continent basement?)
Early Miocene)
Upper Cretaceous-Eocene Belait, Liang and Setap
+ Neogene Melinau Fm
tectonized ophiolite with Shale Formations
igneous rocks (Eocene-Oligocene)
remobilized serpentinite (Miocene-Holocene)
Fig. 25. Regional tectonic map of NW Borneo (modified from Sandal, 1996; Morley et al., 2008; King et al., in press).
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 65

Seafloor high
Oligocene-Miocene Early Miocene
Setap Shale Fmn. Meligan Fmn.
and Temburong Fmn.

D
ee
XI Melinau Limestone Eocene Mulu and

pw
(late Eocene-

at
Kelalan Formations

er
IX Early Miocene).

fo
Approximate

ld
position of deep

an
d
regional unconformity

th
ru
0 25 km

st
be
Shelf edge

lt
N

4
ne
Zo Champion Delta
Frigate Labuan
Province
Fault Island
6 km+
depocentre
3
ne
Zo

SABAH
BRUNEI

SARAWAK
n e2
Zo
Baram Delta
Province

2
ne
Zo

Belait Zone 1
anticline

Belait
syncline

Mt. Mulu

Fig. 26. Map of the structural zones of Brunei, partly based on Sandal (1996), Morley et al. (2003).

comprises present-day Borneo (e.g. Levell, 1987; Hutchison et al., profound effect on the Sabah margin and induced gravity-driven folding
2000). Middle Miocene–Late Miocene shortening has been largely and thrusting in the DWFTB. Alternatively the granite could be a product
confined to the onshore area, with episodic inversion of offshore of slab breakoff (e.g. Von Blanckenburg and Davies, 1995), which would
structures occurring along the NW Borneo shelf of Brunei and Sabah also explain the late-stage uplift of Borneo (e.g. Hutchison et al., 2000;
(Levell, 1987; Hutchison et al., 2000; Morley et al., 2003). The Morley and Back, 2008).
progressive uplift and erosion of NW Borneo from the latest Early The Baram Delta Province built out onto the NW Borneo margin
Miocene onwards resulted in the rapid deposition of thick deltaic from the latest early Miocene to the present day (e.g. Sandal, 1996;
sequences up to 10 km thick around the margin of Borneo, notably the Hall and Nichols, 2002; Morley and Back, 2008). The primary source of
West Luconia, Baram, Sandakan, Tarakan and Mahakam deltas (e.g. the deltaic sediments is the hilly spine of NW Borneo. The region is
Hall and Nichols, 2002; Fig. 4). largely composed of the Sapulut (Late Cretaceous–early Cenozoic),
The most prominent manifestation of Miocene orogenesis in Sabah is Trusmadi and Crocker (Cenozoic) Formations, which are mostly
Mount Kinabalu, which is formed from a Late Miocene granodiorite deepwater sediments, probably of forearc origin (Hutchison, 1996;
sheet (Cottam et al., 2008, 2010). The summit rises up to 4100 m, and van Hattum et al., 2003). The Crocker Formation sands were
was glaciated during the Pliocene. The granodiorite was crystallized predominantly sourced from Cretaceous granites located in southern
between 7.9 and 8.2 Ma and rapidly exhumed by ~4–8 km (Hutchison Borneo and the Sunda Shelf, although a minor ophiolite-derived
et al., 2000; Cottam et al., 2008). The granite is interpreted as a product component is also present (van Hattum et al., 2006). The Sapulut and
of collision-related thickening and its exhumation could indicate deep Trusmadi Formations are compositionally more mature (quartzose
lithospheric processes, such as loss of a lithospheric mantle root (Cottam recycled) (van Hattum et al., 2003). It is important to the hydrocarbon
et al., 2008; Hall, 2009). These authors suggest uplift would have had a potential of the region that these formations provided quartz-rich
66 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

(albeit predominantly fine-grained) sands to the Baram Deltaic impact of deep-seated stress affecting NW Borneo (Morley et al.,
Province (Sandal, 1996). 2003; Tingay et al., 2005).
The NW Borneo margin can be interpreted both in terms of a The structural style of the NW Borneo DWFTB is predominantly
tectonically active margin and gravity tectonics (e.g. Franke et al., landward dipping imbricate thrusts, with hangingwall folds (Fig. 27).
2008; Morley et al., 2008), particularly in the Baram Delta Province The imbricate thrusts lie within a critical taper wedge with a surface
(Brunei, SW Sabah, and the northeastern Sarawak). The onshore slope that ranges between 1° and 2.5° and a basal detachment that
Baram Delta Province is characterized by basement-involved defor- dips 2° to 5° (Morley, 2007a). The wedge thickens from about 3 km
mation that affects the sediments of the early (Middle Miocene) stage near the thrust front to 10 km near the shelf-slope break. The most
of deltaic deposition. Seismic reflection data from onshore and in the external folds exhibit classic imbricate fault splays off the basal
near offshore area shows well developed folds related to inverted detachment, however, the fault geometry at depth becomes unclear
normal faults (e.g. Levell, 1987; Sandal, 1996; Morley et al., 2003, passing into the thicker part of the wedge the fault geometry (Cullen,
2008; Fig. 26). These folds account for up to 3 km shortening as 2010). The folds have a strong expression at the sea floor and
calculated by Hesse et al. (2008). This proximal zone of inversion is influence gravity-driven sediment pathways down the slope (McGilv-
not seen in the shallow gravity-driven systems, and reflects the ery and Cook, 2003; Morley, 2009a; Hesse et al., 2010). The results of

NW SE

South China Sea Sabah Shelf Kinabalu Mt. Kinabalu Telupid Neogene Basin
Deepwater Culmination High
B Fold Belt
Offshore Onshore
++ +
0 +
+ + Forearc
10 UC + ++
Crust
20 Eurasian Plate +
+ +
+
+
+
LC +
++ +
30 + +
Flow of lower crust?
40
Mantle Suture zone and
50 Ductile wedge of Eurasian Plate
detachment zone
km crustal rocks Oceanic slab broken off?
A Pliocene-Holocene sediments West Crocker Formation + +
Kinabalu Granite, and migmaties
Miocene sediments above the
East Crocker/Rajang Formation Miocene melange
Deep Regional Unconformity

Deepwater fold and thrust belt Region of shale diapirs and mini basins,
NW diapirs squeezed by later compression (?) SE
0
TWT (s.)

Lower Plate of Dangerous Grounds


8 continental crust 5 km
Possible mobile shale zone
B (poorly imaged region on
Lower pre-kinematic sequence
(Miocene?)
Upper pre-kinematic sequence
(Miocene?)
seismic data)
Early syn-kinematic sequence, Mid syn-kinematic sequence, Late syn-kinematic sequence,
Late Miocene (?) Pliocene (?) Pliocene - Holocene

NW SE
C
500 ms
10 km

NW SE
D 1 km
0 10 km

Fig. 27. Structural cross-sections across NW Borneo. A) Regional crustal scale transect through Sabah partly based on Hall and Wilson (2000), B) Detailed cross-sction through Sabah
DWFTB, based on 2D BGR seismic line illustrated in Hesse et al. (2008); see inset in transect. C) Cross-section across Brunei DWFTB based on 3D seismic line illustrated in McGilvery
and Cook (2003).
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 67

piston coring in areas of mud volcanoes indicate the larger ones are Garciacaro et al., (in press) provides an excellent overview of the
expelling hot fluids, which contain thermogenic hydrocarbons deepwater setting of this basin.
(Zielinski et al., 2007; Warren et al., in press; Fig. 9). These data The Trinidad–Barbados region formed as a result of subduction of
support seismic reflection surveys that indicates the mud volcanoes the North and South American plates beneath the Caribbean Plate
are sourced from deep, overpressured fluids, and travel along thrust since the middle Eocene (Pindell and Kennan, 2007). Eastwards
faults before breaking vertically to the surface in the cores of movement of the Caribbean Plate with respect to South American
anticlines (Fig. 8; Morley, 2009a,b). The Sabah margin has proven to presently occurs at a rate of ~ 20 mm/yr along much of their plate
be a successful DWFTB play including the following discoveries: Kikeh boundary (Perez et al., 2001; Weber et al., 2001; Saleh et al., 2004).
(~1300 m, 400–700 million barrels of oil reserves), Gumusut-Kakap The El Pilar Fault and associated strike-slip faults are the focus of
(~1300 m water depth, 2.2 trillion cubic feet (TCF) of gas reserve), eastwards translation of the Caribbean Plate with respect to South
Limbayong (0.8 TCF gas reserve, 1000 m water depth), Malikai America (Fig. 28). The strike-slip faults pass into the complex,
(~565 m water depth) Kebabangan and Kamunsu fields (e.g. Ingram extension to transtension-dominated province of the Columbus
et al., 2004; Milton, 2006). Basin east of Trinidad. Evolution of the Caribbean–South America
The Baram Delta province shows well-developed extensional collision zone of eastern Venezuela–Trinidad–Barbados appears to
growth faults in many places. However, the zone of currently active have changed from a more transpressional structural style prior to
extension is narrow (20–30 km), lacks well developed shale “diapirs”, 10–12 Ma to a transcurrent style, which has continued to the present
and lies between a region of folds associated with inverted growth day (Pindell and Kennan, 2007).
faults in the near shore and onshore area, and the deepwater fold and Trinidad is well known for its offshore hydrocarbon fields, and is
thrust belt (Fig. 26). This arrangement of inverted zone-extensional extensively affected by both thrusting and folding, and normal faults
zone-DWFTB is also seen in the modern stress pattern, where the (Wood, 2000; Escalona et al., 2008; Fig. 28). Yet exploration in the
onshore and inner shelf display maximum horizontal stress (Shmax) DWFTB has so far not achieved commercial success (Rajnauth et al.,
orientations approximately sub-perpendicular to the coastline. The 2004). Overpressured mud pipes and mud volcanoes are widespread
maximum horizontal stress directions for the outer shelf and upper and the predominantly thermogenic gas and clasts found in the fluids
slope (zone of active extensional faulting) rotate to lie sub-parallel to from mud volcanoes indicates a deep, overpressured shale origin,
the coastline, then in the DWFTB area, Shmax orientations revert to most likely in Cretaceous and/or Paleogene formations (Deville et al.,
sub-perpendicular to the coast (e.g. Tingay et al., 2005; King et al., 2003). The Columbus Basin displays a number of similar elements to
2009; Fig. 25). the Niger Delta. It is in large-parts a gravity-driven system associated
A newly identified mobile shale-associated region (Shale Prov- with rapid loading of overpressured marine shales by the Orinoco
ince) in northern offshore Sabah displays approximately N–S Shmax delta. It displays large growth fault systems and shale ridges at the toe
orientations (King et al., in press; Figs. 25 and 27). This province is a of the delta on the slope (Wood, 2000; Wood and Mize-Spansky,
region of Late Miocene–Pliocene mobile shale deformation charac- 2009; Fig. 28). The details of how the near-field and far-field stresses
terized by “diapirs” and synclinal mini-basins, that were later interact in Trinidad remain to be assessed. They have the potential to
squeezed and folded (Fig. 27B). King et al. (in press) suggest the be complex since both gravity-driven deformation and far-field
clockwise rotation of stresses within this province with respect to stress-related major strike-slip faults and thrusts affect the area
wells in adjacent areas can be explained by contrasting geomecha- (Truempy et al., 2004; Garciacaro et al., 2011).
nical properties of a soft, thickened mobile shale and surrounding
stiffer sediments of the DWFTB. 4.2.2. Type 2bi weakly linked/unlinked DWFTBs
An effect of the exhumation of Mt. Kinabalu would have been the
latest Miocene and Pliocene transport of large amounts of sediment 4.2.2.1. Makassar Straits. The Makassar Straits lie between the islands
onto the Sabah margin. Figure 25 shows that the Shale Province wraps of Borneo and Sulawesi. West of Sulawesi an extensive deepwater fold
around the northern end of the Kinabalu Culmination and deposition and thrust belt is present, in an area whose geological setting
within the province coincides with the exhumation of Mt. Kinabalu. has proven controversial. North of the Makassar Straits lies the
Hence differential loading of the DWFTB may have been caused by Celebes Sea, which is floored by oceanic crust generated during an
rapid deposition in the Shale Province as sediment supply increased Eocene phase of seafloor spreading (see review in Hall, 2002). While
due to uplift of Mt Kinabalu. there is general agreement that the southern Makassar Straits are
Total shortening across the NW Borneo DWFTB is ~8–13 km. The underlain by continental crust, determining how far oceanic crust
extent of the DWFTB and amount of shortening does not appear to be formation propagated southward into the northern Makassar Straits
strongly related to the distribution of extension on the shelf (Hesse has proved controversial (e.g. Bergman et al., 1996; Hall, 2002). The
et al., 2008; Morley et al., 2008). Line-length shortening (which may latest review of the data indicates that the northern Makassar Straits
considerably underestimate shortening by compaction) is generally also are underlain by highly thinned continental crust (Hall et al.,
equal to, or greater than extension on the shelf (maximum ~ 6 km 2009; Fig. 5).
extension opposite 6 km contraction for the late Pliocene-Recent; The Cenozoic tectonic history of Sulawesi is also controversial. A
Hesse et al. 2008; King et al., 2010; Cullen, 2010). To the NE from very brief outline of the tectonics of this highly complex area is given
offshore Brunei to offshore Sabah the discrepancy between the as background to the setting of the Makassar Straits. Lying on the
amounts of shortening in the DWFTB compared with extension on the eastern margin of the continental core of SE Asia, the island has been
shelf increases (Hesse et al., 2008). The discrepancies between subject to a succession of orogenic events which include the accretion
shortening and extension have been interpreted as demonstrating a of continental fragments derived from Australia, subduction, intense
tectonic component to past and ongoing deformation in the DWFTB volcanic activity, rifting and ophiolite obduction during the Oligo-
(e.g. Hesse et al., 2008; King et al., 2010). cene–Miocene (see reviews in Bergman et al., 1996; Hall, 2002).
Central and eastern Sulawesi is home to the world's third largest
ophiolite (East Sulawesi Ophiolite; Fig. 5), which is a 15 km thick slab
4.2.1.2. Trinidad (deepwater Columbus Basin). The DWFTB in Trinidad of oceanic plateau crust of possible Cretaceous origin (Kadarusman
is not well described in the literature, but because examples of mixed et al., 2004). According to Parkinson (1998) obduction occurred
near- and far-field stress systems are few, as are examples of DWFTBs during the Late Oligocene. However, there are other ophiolite
associated with strike-slip margins, the deepwater Columbus Basin is fragments present on the island with different geochemistry, hence
briefly discussed in this section to help fill this gap. The paper by the Sulawesi ophiolites are likely to be composite and were emplaced
68 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

0 50 100 km
St Lucia
A N

Mud volcanoes Thrust faults


GRENADA
BASIN

Normal faults Strike-slip faults

s
ine
Barbados 13°00’
ad
en

ASIN
Gr

T
B

ES
AGO

CR
TOB

S
A DO
RB
BA
Grenada

GE
12°00’

RID
SIN

S
BA

DO
OS

A
RB
AD

BA
RB
BA
A
Tobago

11°00’

Trinidad

El Pilar Fault

A’ 10°00’

COLUMBUS
BASIN

ORINOCO
DELTA

62° 00’ 61° 00’ 60° 00’ 59° 00’ 58° 00’

A Northern basin Darien Ridge Columbus Foreland Basin A’


NW Fold-thrust belt SE
0
Seafloor
Two-way travel time (s.)

2
+

4
+ +
6
?
8 ?

10
20 km
12
B
Quaternary Neogene Palaeogene Cretaceous

Fig. 28. A) Regional tectonic map of the Trinidad, Columbus Basin, Barbados area (compiled from Deville et al., 2006). B) Line drawing of interpretation from seismic reflection data,
across the western part of the Barbados accretionary prism, modified from Garciacaro et al., (2011), see A) for location.

during more than one tectonic event (Hall, 2002). Obduction of the 2002). Deformation of the DWFTB in the Makassar Straits dates from
East Sulawesi Ophiolite is commonly assumed to have been followed the Early Pliocene (Calvert and Hall, 2007).
by Late Oligocene–Early Miocene continent–continent collision of an The DWFTB is part of a dominantly east-vergent system of folds
Australian craton-derived block (eastern Sulawesi) with western and thrusts that affect western onshore Sulawesi and the eastern
Sulawesi, which resulted in fold and thrust belt development and Makassar Straits. This DWFTB is a system of detached folds and thrusts
extensive lithospheric melting-related magmatism (see reviews in (Fig. 5; Bergman et al., 1996; Calvert and Hall, 2007). However,
Bergman et al., 1996, and Hall, 2002). Marine sedimentation beyond the thin-skinned thrust front, and beneath the thin-skinned
continued uninterrupted by contractional deformation in western folds, seismic reflection data shows evidence for inversion of older rift
Sulawesi and the Makassar Straits throughout the Miocene (Hall, basins, and some seismic lines indicate that the DWFTB may be linked
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 69

Fig. 29. Comparison of topography across the Baram and Mahakam deltas, and west Sulawesi.

with, or affected by movement on basement-involved fault systems. The uplifted, predominantly ophiolitic and volcanic-sourced fold
The timing of deformation, and deep-seated nature of the stresses and thrust belt to the east sourced Neogene sediments within the
affecting the DWFTB deformation in the Makassar Straits is similar to DWFTB. Several exploration wells have been drilled in the shallow
DWFTB deformation in NW Borneo, and in also the Mahakam Delta water and onshore equivalent section of the DWFTB. The potential
area. reservoir rocks are fine-grained sandstones, with a large basic igneous
Most of the thrusting and folding in the Mahakam Delta is located component. Diagenetic alteration of the igneous-derived clasts to
onshore or on the shelf, not in deepwater (Bergman et al., 1996; clays and the fine-grained nature of the deposits resulted in a low-
Ferguson and McClay, 1997), but is a manifestation of the regional far- permeability, low-effective porosity reservoir. Hence no good quality
field stresses that affect both the onshore areas and the DWFTBs. The reservoir rock has been established to date for the DWFTB system. The
well established NW–SE present-day maximum horizontal stress apparent lack of reservoir is the most significant risk to the presence
direction in the Mahakam delta area fits with a heavy influence from of an economic petroleum system associated with this DWFTB.
far-field stresses (Tingay et al., 2010, in press) The West Sulawesi Another DWFTB complex lies north of the Banggai-Sula micro-
DWFTB in the Makassar Straits is not associated with a significant continent, between Sulawesi and the Banda Sea(Fig. 3). This
delta. The western margin of Sulawesi plunges abruptly into microcontinent probably collided with the eastern ophiolite and
deepwater, with virtually no shelf present, unlike the well-developed northern volcanic arms of Sulawesi in the Early Miocene (Ferdian et
shelf of NW Borneo (Fig. 29). Hence, there is no ambiguity regarding al., 2010). The area also underwent Early Pliocene and south-directed,
the nature of the driving stress (near- or far-field) in the Makassar E–W trending dextral transpressional deformation that formed a
Straits as discussed above for NW Borneo (Baram Delta Province and DWFTB as the Molucca Sea Collision Complex (Fig. 3) impinged on the
Sabah Margin, NW Borneo). The absence of a delta and landward- Sula Platform (Ferdian et al., 2010).
dipping basal detachment indicates there is no component of shallow
gravity driven deformation. 4.2.2.2. Banda Arc. The Banda Arc lies in eastern Indonesia and Timor-
The absence of a shelf today offshore western Sulawesi is also in Leste (Fig. 24). The region provides an example of the early stages of
strong contrast with the older Cenozoic history of the basin. Following continental margin-arc collision (Karig et al., 1987; see review in Hall
Eocene rifting, the Makassar Straits underwent post rift subsidence and Wilson, 2000; Standley and Harris, 2009). The Banda Arc has a
(with extensional activity on some faults continuing into the prolonged and complex history of Cenozoic deformation, with
Oligocene; Calvert and Hall, 2007, Hall et al., 2009). The centre of Standley and Harris (2009) reporting 5–6 deformation phases of
the Makassar Straits is characterized by deepwater sedimentation alternating extension and compression, that are associated with
during the Neogene. However, Calvert and Hall (2007) describe predominantly top to the SE collisional deformation and top to the
extensive shallow marine conditions as being present throughout the south to SE extension. Deformation prior to about 25 Ma was largely
Early Miocene onshore in NW Sulawesi, and persisting in places until related to subduction of the Indian Ocean lithosphere at the Sunda
the Middle or Late Miocene. This former shelf area has been uplifted (Java) and Celebes Trenches. Subsequently, events testifying to
by folding and thrusting, which was sufficiently young and rapid that continued Australia–SE Asia convergence are the collision of the
a new, broad shelf has yet to form. Sula Spur with the North Sulawesi volcanic arc, ophiolite obduction in

EURASIAN PLATE WETAR STRAIT TIMOR TIMOR TROUGH AUSTRALIAN PLATE


INNER BANDA ARC BANDA ALLOCHTHON ONNER BANDA ARC
Permo-Triassic distal
Australian margin sediments
Forearc crust Post-rift slope and Syn-orogenic deposits Post-rift shelf deposits
NNW rise deposits SSE
0

+ + + + + Permo-Triassic +
W Australian crystalline basement
SU ETA +
+ + + + pre-rift sediments
TU R + +
+ + + +
Kilometres

RE +
WET +
AR T + +
H RUS
T +
Cambrian-Carboniferous
50
pre-rift sediments
Asian mantle
Australian mantle

Australian
oceanic lithosphere
100

Fig. 30. Regional crustal-scale transect, illustrating the early-stage collision of the Australian Plate with the Timor island arc. Redrawn from Hall and Wilson (2000).
70 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

South Sulawesi, and broad counter-clockwise rotation of the Borneo– Extensional basins ranging in age from Permian to Paleogene may
Java region of Sundaland (Hall, in press). At about 15 Ma, the oceanic contain excellent Mesozoic source rocks (Charlton, 2004). Hydrocar-
Banda back-arc basin began to form as a result of subduction rollback bon prospects are available in shallow marine to onshore environ-
(see review by Hall, in press). Thinned crust of the Australian ments in these inverted rift basins, as demonstrated by the Oseil Field
continental margin was subducted beneath the Banda Arc, whilst the in Seram, where the Upper Triassic–Lower Jurassic Manusela
attached oceanic lithosphere was broken off during the late Neogene Limestone reservoir is sealed by the external, thin-skinned thrust
(Hall and Wilson, 2000; Fig. 30). The subduction channel may have and mélange belt (Nilandaroe et al., 2002). In such a setting, large and
become clogged by Australian margin crust only about 3 my ago relatively simple structures may occur beneath the structurally
(Standley and Harris, 2009). The deepwater troughs fringing the complex, thin-skinned fold and thrust belt (Charlton, 2004; Pairault
Banda Arc such as the Timor Trough have evolved from a subduction- et al., 2003; Sapin et al., 2009; Fig. 31) The most external and youngest
trench setting to a collisional flexural (deepwater) foreland basin structures of the Timor, Tanimbar and Seram fold-and-thrust belt
(Carter et al., 1976; Audley-Charles, 1986; Ziegler et al., 1998). (Fig. 24) are relatively simple hangingwall anticlines (Fig. 31), located
Notably the greatest water depths are found not in the sites of the in water depths between 1500 and 2000+ m. However, in places,
former accretionary prisms, but in the back arc location (Fig. 24). such as the Onin and Kumawa domes east of Seram (Fig. 32) and the
The upper plate of the collision zone includes the islands of Timor, Kei Islands compressional reactivation of deeper seated Mesozoic
Seram, Tanimbar and Kei. Hydrocarbon seeps on a number of islands, extensional fault systems caused late stage uplift of the deepwater
including Timor, have been observed, but only Seram is at present an fold and thrust belt to shallow waters, or subaerially, inducing its
established hydrocarbon province (Charlton, 2004). The broad zones erosion (Fig. 31).
of deformation comprise both thin-skinned thrusting, and thick- Extensive diapirism, forming diapir fields up to 100 km long and
skinned deformation of the basement, including inversion of older 30 km wide, is observed in Australian passive margin sediments
normal faults; the degree of overthrusting is considerable in places. before they enter the Timor accretionary complex (Breen et al., 1986;
For example, Permo-Triassic sedimentary rocks of the distal Austra- Masson et al., 1991). These diapirs were interpreted by Barber and
lian margin observed on Timor have been thrusted onto Mesozoic– Brown (1988) as resulting from overpressured fluids generated
Cenozoic sedimentary sequences of the Timor Trough flexural forearc initially by disequilbrium compaction, due to the overburden of the
foreland basin. In turn the Australian margin sediments that crop out accretionary complex, that were transferred along Jurassic breccias
in Timor have been overridden to the SSE by the Banda Arc forearc. and sandstones marking the break-up unconformity, to mobilize well-
Oceanic lithosphere in west Seram was obducted over an Australian- consolidated Lower Cretaceous clays. The Lower Cretaceous clays
derived microplate of continental crust during the Miocene (Linthout formed diapirs that were incorporated into the accretionary complex
et al., 1996). during subduction. Outcrops in Timor also indicate that diapiric

1) Tortonian SW NE

South-western Bird’s Head Microplate Margin ??


2) Messinian

Seram fold and thrust


belt front

3) Late Pliocene

Misool-Onin-Kumawa Ridge

Ductile deformation ??

4) Present

High Bouger anomaly


on 250 km high-pass filter

10 km

Steenkook Formation Klasafet Formation New Guinea Limestone

Permian-Paleocene shales Brittle/Ductile crust interface Inactive fault Active fault

Fig. 31. Conceptual evolutionary diagram of the Seram fold and thrust belt (redrawn from Sapin et al., 2009). For location see Fig. 32.
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 71

Fig. 32. Tectonic map of the Seram collision zone (based on Sapin et al., 2009), showing location of Fig. 31.

mélange is a very important component of the Neogene accretionary highly unusual example of an active margin. Accretionary prisms form
prism (Barber et al. (1986). This mélange is derived from the Lower a wedge-shaped profile, where new material is accreted to the thin
Cretaceous clays discussed above, and in the north from Permo- end of the wedge. The oldest and thickest part of the accretionary
Triassic clays. On Timor, very large areas are occupied by the Bobonaro prism can be thickened internally and increased in volume by
Scaly Clay that is associated with intrusive features, including forcing underplating of material. Consequently the wedge becomes uplifted
of the clay matrix along internal joints and fractures and the break-up and exposed, to commonly form a chain of islands. These islands, such
of numerous blocks under high overpressures (Barber et al., 1986). as the Andamans, Nicobar, Barbados, Timor, Ramree (Burma), or
This evidence needs to be considered as a counter point to the onshore ranges (e.g. the Makran) may display encouraging signs of
arguments that large, mobile shale masses do not exist. While regular hydrocarbon generation with oil and gas seeps from deepwater
folds and thrusts can occupy parts of accretionary prisms, large areas sediments. The seeps are often from mud volcanoes. The sediments of
may also be broken down into more chaotic units by transfer of huge the islands tend to have undergone multiple episodes of folding and
volumes of overpressured fluids to the distal edge of the accretionary thrusting, shale diapirism, and sometimes a phase of extensional
complex. tectonics (Pindell and Kennan, 2007; Maurin and Rangin, 2009). This
complex deformation can be understood in the context of critical
4.3. Type 2bii active margins taper theory, where out-of-sequence deformation is necessary to keep
thickening the internal part of the wedge so oceanward propagation
4.3.1. Introduction of the thrust front can continue (Platt, 1986). Conversely, thickening
Active margins extend for over 30,000 km around the world, and of the wedge also promotes gravitational collapse and episodes of
their associated accretionary prisms represent a vast accumulation of extensional faulting (Platt, 1986). Such a complex deformation history
sediment. Yet, despite their great volume, their contribution to means that trapping geometries, if they ever exist at all, will tend to be
conventional world petroleum reserves is small, particularly com- small. Typically, the older, deeper parts of the accretionary prism have
pared with the prolific DWFTBs of salt-associated passive margins been exhumed in the islands, hence compaction and diagenesis have
whose total length is about one third of the active margins. There is considerably reduced reservoir porosity and permeability. With the
one exception to this statement; it is the South Caspian Sea, which is a advent of deepwater exploration, there is the possibility that the
72 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

simpler structures further offshore will avoid the disadvantages of the other rivers; Indo-Burma Ranges–Ganges–Bengal Fan, and precursor
proximal part of the wedge, and permit exploration for younger, less systems. A schematic illustration of the general relationships between
deeply buried turbidite reservoirs in larger, simpler structures. large deltas prograding into an active trench system and providing an
Out-of-sequence faulting is also thought to signify the onset of evolving accretionary prism with sand-prone sediments is shown in
significant lithification (i.e. increasing elastic strength) within the Figure 33. The deep-water Makran fold and thrust belt trends normal
accretionary prism, which permits slip to occur along discrete large- to the drainage systems of its mountainous hinterland, while the
scale blocks. The modern activity of out-of-sequence thrusts suggests Indo-Burma Ranges and Trinidad–Barbados accretionary prisms are
the seawardmost out-of-sequence thrust approximately defines the charged along strike by fluvial-deltaic systems which prograde from
seaward edge of the elastically rigid upper plate (Moore et al., 2007). the related onshore foreland basins (Figs. 28 and 37). The ultimate
The following discussion focuses on present day active margin end-member to these settings is the South Caspian Sea, which has
DWFTBs and their hydrocarbon accumulations. Therefore, rather than undergone in excess of 10 km subsidence since the Pliocene due to the
reviewing the vast range of known active margin settings, we center influx of sediments from adjacent orogenic belts, particularly the
on the more atypical settings with the highest hydrocarbon potential. Greater and Lesser Caucasus (Allen et al., 2002; Brunet et al., 2003).
This means classic example of accretionary wedges, which are Accretionary prisms display a great range in width depending
associated with the subduction of large oceanic plates from areas upon their stage of development and sediment supply, which is
such as Japan, Sumatra–Java, and Chile are not discussed herein, since derived both from fluvio-deltaic sediments supplied to the upper
they have a relatively minor conventional hydrocarbon potential plate, and sediments scraped off the lower plate. For example the
(although the widespread occurrence of gas hydrates in accretionary frontal 50 km of the Apennines displays an average cross-sectional
prisms, represents a potentially vast unconventional resource). area of 500 km2, while the Northern Barbados Ridge is only 100 km2
Improved conventional hydrocarbon potential occurs where the (Bigi et al., 2003). This variation in sediment volume resulted in the
more simply deformed external zones of DWFTBs lie in water depths Apennines displaying a deeper detachment and higher elevations
of 3000 m or less, and where large fluvio-deltaic system(s) may have than the Barbados Ridge (Bigi et al., 2003).
provided quartz-rich turbiditic sands to the evolving DWFTB. Areas
that match these criteria are the narrow, complex accretionary prisms 4.3.2. Barbados
of the Caribbean, (e.g. Barbados, offshore Cuba), and those on the The Columbus Basin of Trinidad (Fig. 28) has been briefly
trend of major continent–continent collision zones (e.g. The Makran, described in Section 4.2.1.2. There is a northwards transition from
South Caspian Sea and the Indo-Burma Ranges). Such settings are the Columbus Basin, to the accretionary prism setting of Barbados.
invariably transitional to continent–continent collision zones: the Elements of the Trinidad geological history are applicable passing into
Indo-Burma Ranges grade from an active margin setting northwards the deepwater Barbados accretionary prism. For example the basal
to an early-stage collisional zone. The Barbados accretionary prism overpressured mobile shales, and the hydrocarbon source rocks in
passes laterally into the transpressional setting of Trinidad to the both cases appear to be upper Cretaceous marine shales with La Luna
south and Cuba to the north. While the Makran passes into continent– Formation-affinity (e.g. Hill and Schenk, 2005; Deville et al., 2006).
continent collision zones to the west (Zagros) and east (Himalayas), The Woodbourne Field, onshore Barbados has produced upper
and is associated with the remnant of the Neotethys subduction zone Cretaceous-sourced oil from Cenozoic reservoirs since the 1970's,
(e.g. Stampfli and Borel, 2004; Hafkenscheid et al., 2006). Develop- and has provided encouragement for the new offshore exploration in
ment of major Cenozoic orogenic belts (Andes, Himalayas) was deeper water (Dolan et al., 2004; Hill and Schenk, 2005).
accompanied by the development of large sediment-laden river The Barbados Ridge Complex narrows from 250 km in the south to
systems that debouched into the following deepwater oceanic 100 km in the north, and its thickness varies from 200 m to 7000 m from
domains: Trinidad–Barbados–Orinoco River; Makran–Indus and north to south (Moore et al., 1990). This thickness change reflects the

Onshore/nearshore

Ophiolite and volcanic-arc


Continental terrane
orogen or Shelf
strike-slip belt

Slope/deep marine

Major fluvial
system and delta

Trench

Sediment transport Sediment transport from


Anticline
from arc terrane continental area

Fig. 33. Schematic illustration of an accretionary prism with a dual sedimentary source due to the proximity of the DWFTB to a continental orogenic belt. The situation can occur
when a late-stage accretionary prims lies oblique to a continent-continent collision zone. For example the N–S River Ganges traverses E–W trending Himalayan structures, and
enters the Bay of Bengal and feeds a vast volume of sediment into the N–S trending foredeep basin to the Indo-Burma Ranges. Smaller drainage systems enter the Bay of Begal from
the Indo-Burma Ranges to the east.
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 73

importance of the Orinoco Delta, located to the south of the ridge, as the 4.3.3. The Makran
Barbados Ridge Complex's primary sediment source (Belderson et al., The Makran DWFTB is an accretionary prism about 900 km long,
1984; Faugères et al., 1993; Dolan et al., 2004; Fig. 28). Active thrusting which lies in relatively modest water depths of up to 3000 m and
and folding occurs up to 30–100 km from the toe of the prism (Mascle et grades to the NW into the frontal structures of the Zagros fold and
al., 1990; Huyghe et al., 1999, 2004, Fig. 28B). According to Mascle et al. thrust belt. This accretionary prism developed on a fragment of
(1990) the prism is dominated by a series of northwest-dipping imbricate Cretaceous Semail Oceanic crust (Stampfli and Borel, 2004) that
thrusts, spaced between 2 and 12 km (average spacing ~7 km). survived in the gap between the Arabia–Eurasia collision zone of the
Deformation becomes more intense to the NW, with faults being more Zagros Mountains to the west, and the India–Eurasia collision zone of
closely spaced, and shortening increasing from less than 1 km on the the Himalayas to the east. The Makran accretionary prism developed
external imbricates to up to 5 km on thrusts further to the NW. Fold axial throughout the Cenozoic as the Arabian Plate was subducted under
trace length in the south of the prism ranges between 10 and 20 km long, the Eurasian Plate (Harms et al., 1985), with subduction probably
and increases to 30–40 km further north. The decrease in sediment commencing during the Paleocene–Eocene (Platt et al., 1988; Byrne et
thickness from south to north affects fold and fault style within the prism: al., 1992). The proximity to the India–Eurasian collision zone resulted
relatively low-angle thrusts (~20° dip) deform the thin sequences, while in a large sediment supply to the Makran accretionary prism (Fruehn
faults with higher dip (up to 40°), and better developed piggy-back basins et al., 1997). Large turbidite systems were deposited during the late
characterize the thicker sequences (Mascle et al., 1990). Paleogene to Miocene, incorporated into the accretionary prism, and
The basal detachment of the accretionary prism is strongly subsequently uplifted, eroded and recycled into the younger, ocean-
overpressured, and has flowing fluids with pressures N90% of ward propagating parts of the accretionary prism (Harms et al., 1985;
lithostatic (Housen et al., 1996). Chlorine stable isotope ratios Platt et al., 1985). In this fashion the accretionary complex has
measured from mud volcanoes at the outer edge of the Barbados propagated seawards at an estimated rate of ~10 cm yr− 1 (White,
accretionary prism provide evidence for a fluid reservoir at the base of 1982; Platt et al., 1985). Plate convergence at present is about 2.7–
the decollement containing seawater and water released during gas 4.2 cm yr− 1 (DeMets et al., 1990; Vernant et al., 2004).
hydrate destabilization (Godon et al., 2004). Large fluvio-deltaic systems have fed onto the continental margin,
Hydrocarbon exploration in the offshore area of Barbados began in so that offshore growth faults are imaged on seismic reflection data
1996 with Conoco (now ConocoPhillips; Dolan et al., 2004). The (Grando and McClay, 2007; Fig. 34). These growth faults appear to be
exploration concept required the presence of oil-prone source rocks linked to the toe thrusts. The range of active structures in the Makran
equivalent to the La Luna Formation facies, Neogene reservoir rocks accretionary prism comprises near-shore growth faults, upper slope
(turbidites) derived from a proto-Orinoco sediment source, and to mid-slope shale diapirs or mud pipes and an outer-slope imbricate
accretionary prism folds and thrusts providing the main traps (Dolan zone (Grando and McClay, 2007; Fig. 34), These characteristics are, to
et al., 2004). The exploration effort resulted in the collection of over a degree, akin to those of a large delta-driven Type 1ai system.
36,000 km of 2D seismic data, 3D seismic data, plus an extensive Nevertheless, and as evident from Fig. 34, the amount of shortening
program to identify hydrocarbon seeps by the collection of high- achieved in the distal imbricate zone exceeds by far the amount of
resolution multi-beam swath bathymetry, piston coring and dredge extension along the proximal normal fault systems. Correspondingly
sampling. Upper Cretaceous and Cenozoic source rocks were the Makran system qualifies as a Type 2bii DWFTB.
identified (Dolan et al., 2004). In the older, uplifted, onshore part of the Makran system satellite
Seismic reflection data revealed that multiple episodes of images show a large-scale example of a Miocene growth fault
compression, followed by extension and collapse, had affected the controlled depocentre that has been subsequently folded (Fig. 35).
accretionary prism. Mud pipe intrusions and mud volcanoes occur In this regard the Makran resembles the Baram Delta Province. The
extensively (Fig. 28). Conjugate (NW–SE and NE–SW trending) strike- growth fault depocentre in Figure 35 is of a similar size to major
slip fault systems consistent with E–W compression are also present. growth fault depocentres in the Baram and Niger Deltas.
Many folds trend NE–SW and are bounded by NW–SE transpressional Development of major normal faults in accretionary prisms has
faults. According to Dolan et al. (2004) seal integrity is regarded as a been explained by Platt (1986) and Grando and McClay (2007) in
major challenge due to pervasive shale diapirism and young faulting. terms of a critical wedge model (Fig. 36A). Thickening of the back end
Migration of overpressured fluids is commonly linked with fault of the wedge (for example by underplating) can trigger extensional
systems. The first offshore exploration well (Sandy Lane-1) estab- collapse of the wedge either by differential loading or gravity sliding.
lished the presence of thick, well-developed, sand-rich turbidite lobes, This model implies extension occurs regardless of the nature of
with biogenic and thermogenic gas saturations up to 15% recorded in sedimentation affecting the margin. However, the growth fault
some sands. However, the structure is thought to have been breached example from the Makran (Fig. 35) indicates a large delta was
by Late Pleistocene strike-slip fault activity. NW–SE strike-slip fault present during the Miocene, and differential loading probably
trends appear to be the most significant hydrocarbon migration triggered the development of the large depocentre. Like in NW
pathways, and the main cause of trap breach (Dolan et al. (2004). Borneo, the normal faults could have linked with compression down-

N S
0

Imbricate zone 2
TWT (s.)

Deformation Front
4

6
10 km
8
Chaotic zone
Himalayan Turbidites Makran Sands Growth I Growth II Growth III on seismic

Fig. 34. Cross-section through the Makran DWFTB (Grando and McClay, 2007). The chaotic zones on the seismic data could be indicative of shale diapirs, narrower shale pipes, or gas
chimneys.
74 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

Fig. 35. Satellite image of a large extensional growth-fault bounded depocentre in the southern Makran, Pakistan. This depocentre has been folded and is now involved in a syncline.

Slope too high: extension Shale diapirs or pipes Slope too low: contraction
A
0
km
20
low rate of frontal accretion

40
large-scale
underplating
50 km
Active growth fault depocentre
High sediment supply - low slope Shale diapir or pipe Slope too low: contraction
B Inverted/folded growth fault depocentres
0
km

20
Deformation in accretionary prism
driven by both near-field stress
and far-field stress
40 large-scale
underplating
Folded inactive extensional detachment

Fig. 36. Models for the development of extensional faults in an accretionary prism setting. A) Extensional collapse in response to overthickening of the inner part of a critical taper
wedge, (Platt, 1986). B) Extension due to differential loading of the inner wedge by a large delta. Older extensional fault depocentres have become inverted.
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 75

dip (Fig. 36B), and contributed to shortening within the wedge, then
was subsequently inverted. Such features are seen in both the A
Main Boundary Thrust
Pakistan (Fig. 35) and Iranian Makran (Hosseini-Barzi and Talbot,
2003).
To date, only four wells have been drilled in the offshore Makran, Shillong Plateau

Ganges R.
some of which had gas shows, but apparently established no
commercial accumulations. Numerous gas seeps associated with 25°N
mud diapirs are located near the coast (as reviewed in Khan et al.,
C
2007). Temperatures within the accretionary prism are relatively low, 9
with geothermal gradients averaging about 2 °C/100 m, and average
heat flow is 47 mWm− 2, which suggests maturation of hydrocarbons

Kabaw Fault
between depths of 3000 and 5000 m (Ahmad, 1969; Kaul et al., 2000;

Sagaing Fault
Kaladan Faul
Khan et al., 2007). Several potential source horizons are indicated
Shan
including: the Upper Miocene Parkini Shales, kerogen type II–III, TOC Plateau
0.5–1% (Khan et al., 2007); and the Middle Miocene (abyssal-slope)

Cr tal
Panjgur Shales, type II–III, TOC 0.7–5.6%. The offshore Parkini turbidite

t
nic en
us
ea ntin
sandstones of Miocene–Pliocene age appear to offer reservoir

t
Oc Co
potential, particularly the Panjgur Sandstones (Platt et al., 1985; B
Khan et al., 2007). However, there are indications of ophiolitic and
volcanic contributions from the sediment source area in the 12 20°N
quartzolithic Panjgur Sandstones, which gives rise to a complex
diagenetic history with secondary porosity creation and destruction
and formation of a range of cements (chlorite, ferroan calcite, ankerite
and dolomite; Garzanti et al., 1996; Grigsby et al., 2008). Section B
A
4.3.4. Andaman subduction zone–Indo-Burma Ranges
The N–S trending Indo-Burma Ranges lie in western Myanmar and Yangon
18 Section C
eastern Bangladesh (Fig. 37). Onshore they are composed primarily of
Palaeogene deepwater sediments, and Cretaceous mélanges, contain-
Microplate
Indo-Burma
ing blocks of gabbro, pillow basalt, serpentinite, banded chert, Fold and Thrust Belt
Burma

limestone and schist, interpreted as having developed as result of N Myanmar Central


to NE directed subduction during the Mesozoic–Cenozoic (Mitchell, Basins and Pegu
1993). The western, external parts of the Indo-Burma Ranges consists Yoma
of folded Miocene–Pliocene fluvio-deltaic sediments, essentially India/Eurasia 15°N
sourced by the river Ganges, which were folded during the Boundary Coco Island
ch

Pliocene–Holocene (Sikder and Alam, 2003; Maurin and Rangin,


n

Fold axis
Tre

2009).
an

Fault Andaman Basin


The Indo-Burma Ranges initially developed as a result of
dam

subduction of the Neo-Tethys Ocean, and then, from the Senonian 90°E 95°E
An

onward, of the Spongtang Ocean beneath the active southern margin NW WellA4 SE
of Eurasia (Stampfli and Borel, 2004; Hafkenscheid et al., 2006; Hall et 0
Gravity slide P
1 M
al., 2009). The Indo-Burma Ranges are, in part, an old accretionary O
2 E
prism complex that developed during the Late Cretaceous and
TWT (s)

Paleogene and show a complex history of deformation, ophiolite 3


emplacement, uplift and erosion (see: Mitchell, 1993; Acharyya, 4 Mud diapir
2007). The coupling of India with NW Myanmar, and the collision of 5 B
India with Eurasia terminated the early accretionary history during 6 P = top Pliocene O = top Oligocene
10 km
M = top Miocene E = top Eocene
the late Paleogene (Searle and Morley, in press). However, earthquake 7
SW NE
activity from the Burma seismic zone indicates that subduction 0
Wedge
processes are still active beneath western Myanmar and Bangladesh 1
(Stork et al., 2008). This section focuses on the post India–Eurasia 2
coupling history of the Indo-Burma Ranges.
3 Mud diapir?
India has undergone highly oblique convergence with continental
4
core of SE Asia during the Cenozoic. GPS measurements indicate India 1 km C
is presently moving about 35 mm/yr-1 northwards with respect to 5
Sundaland. This motion is accommodated by distributed deformation Fig. 37. A) Tectonic map of the Indo-Burma Ranges showing location of structural cross-
on numerous faults across the Burma microplate,of which the Sagaing sections given in B and C. (compiled from Maurin and Rangin (2009) and Nielsen et al.
Fault, with a dextral displacement of b18–20 mm/yr− 1, is the largest (2004)). Arrows (A,B,C) and numbers indicate the direction and relative convergence
(Fig. 37; Vigny et al., 2003). rate of India with respect to the upper plate margin (all rates in mm/yr), which
progressive rotate from 035°/18 mm/yr (south Myanmar) to E–W/9 mm/yr
The outer Indo-Burmese wedge forms the western margin of the
(Bangladesh fold and thrust belt), from Nielsen et al. (2004).
Burma microplate and lies predominantly on oceanic crust (Curiale et
al., 2002). It underwent folding from the late Miocene–Recent
(Maurin and Rangin, 2009). The wedge is characterised by a broad occurred in the north above an efficient Oligocene deep marine,
region of folding in the north, covering the onshore and offshore areas overpressured detachment and covers a broad region up to 150 km
of Bangladesh and Myanmar. The fold and thrust belt narrows to the wide (Sikder and Alam, 2003; Maurin and Rangin, 2009). Five
south and becomes increasingly focussed offshore. Folding has hundred kilometres further south, the young fold and thrust belt is
76 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

just 20 km wide and is composed of just 2–3 folds along a narrow conditions during the Pliocene. The deepwater gas discoveries in
slope, that are bounded by a major transpressional fault on the NW Myanmar are stratigraphically trapped in Pliocene turbidites and
western margin (Fig. 37A,B). Young deformation also affects the shelf indicate that a viable deepwater petroleum system exists in the area.
area, with the development of normal faults, and mud diapirs; However, the timing of folding in this area is young, during the last
subsequently the entire region underwent extensive inversion 2 Ma (Maurin and Rangin, 2009). A combination of uplift due to
(Nielsen et al., 2004; Maurin and Rangin, 2009). folding, and infilling of accommodation space by rapid north to south
The change in structural styles from south to north indicates that progradation of the Bengal Fan, together with lesser input from the
the highly distributed oblique deformation of the Burma microplate Indo-Burma Ranges to the east (Allen et al., 2008) has resulted in the
becomes increasingly focussed towards the south on the oblique- northern part of the Outer Burma Wedge rapidly developing into a
dextral Andaman Subduction Zone. In the southern region (near the shallow marine environment in the relatively recent geological past.
north Andaman Islands), the 3.5 cm/year motion of India relative to At present, the northern 550 km of the Outer Burma Wedge lies either
Sundaland is taken up, in equal amounts, by the Andaman Subduction in shallow water or onshore. This rapid progradation means that, in
zone and Sagaing–West Andaman Fault system. However, this motion places, plays for deepwater Late Miocene or Pliocene turbidites folded
is fully partitioned across the whole Burma microplate in the north by accretionary prism-related deformation can be found in a shallow
(Nielsen et al., 2004). Deepwater troughs are present both east and marine setting. Yet these plays have not met with economic success
west of the Coco Islands. Considerable folding and thrusting has offshore so far. One problem may be that folding is too young, with
inverted deepwater basins on the east side of the islands. The respect to hydrocarbon migration.
occurrence of a deepwater basin on the east side is probably due to the
presence of the Andaman Sea Spreading Centre to the south that was a 4.3.5. South Caspian Sea
source of recent subsidence in the region, coupled with the sediment- The South Caspian Sea is a prolific oil province with reserves likely
starved nature of the area. to be in excess of 30 billion barrels of oil equivalent (BBOE). The South
The Bay of Bengal is well known for the massive accumulation of Caspian Basin extends 680 km N–S and 500 km E–W, with maximum
Cenozoic deltaic sediment associated with the Bengal Fan, which water depth of ~1000 m (Fig. 38). The South Caspian Sea is thought to
exceeds 20 km near the Bangladesh coastline (e.g. Curray and Moore, be underlain by about 10 km thick oceanic crust, which is covered by
1971; Metevier et al., 1999; Curiale et al., 2002; Clift, 2006). Much of up to 26–28 km of Mesozoic and Cenozoic sediments (Berberian,
the northern fold belt area in Bangladesh accumulated sediments in a 1983; Jackson et al., 2002; Knapp et al., 2004; Fig. 39A). This crust
deep marine environment during the Early Miocene (Gani and Alam, originated during the Mid-Late Jurassic opening of the Izmir–Ankara–
2003). By the Late Miocene, the area was undergoing deposition by South Caspian back-arc basin (Golonka, 2004; Stampfli and Borel,
braided fluvial systems, and folding developed in a continental 2004).
environment (Sikder and Alam, 2003). Much of the Outer Burma The South Caspian Basin is bounded to the north by the Apsheron
wedge along the present coast of Myanmar lay in deepwater ridge that links the Great Caucasus and Great Balkhan –Kopet-Dagh,

Gre
ate 48° 52° 56°
rC
auc
asu
s
-400 m water depth
-4
00
m

fold
North Caspian Sea Basin
large thrust or strike-slip fault

Kura Basin
Ap
40° she
ron
Rid
ge

Talysh

South Caspian Sea Basin Kopet Dagh

-4
00
37° m

100 km Alborz Mtns.

Fig. 38. Regional tectonic map of the Southern Caspian Sea and surrounding areas. (modified after Allen et al., 2002).
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 77

South Caspian North Caspian


A Caspian Sea Mud volcano

Apsheron
Ridge
Maykop Fm. Continental
Mesozoic sediments crust
Line of section
Oceanic (?) crust
South Caspian
Sea
Folded and thrusted complex

South North
0
B South Caspian Sea

Productive Series

Productive Series
Depth (km)

Maykop Series
10 North
Caspian
Mesozoic-Eocene
Crust

15

Oceanic (?) crust


20 km
20

Fig. 39. A) Crustal-scale model for the tectonic setting of the Southern Caspian Sea and Apsheron Ridge. B) Structural cross-section based on 2D deep reflection seismic data,
illustrating structural style of the boundary zone between Absheron Ridge and Southern Caspian Sea (after Stewart and Davies, 2006).

which evolved by Late Eocene and younger inversion of rifted basins. Fig. 39A). This slab shows seismicity at depths of 30–80 km and is
The South Caspian Basin is bounded by the Alborz fold and thrust described as incipient subduction (Jackson et al., 2002; Knapp et al.,
belt to the south, which remained active during Neogene times 2004).
(Allen et al., 2002; Jackson et al., 2002; Knapp et al., 2004). During The close proximity of the South Caspian Basin to continent–
the complex late Mesozoic and Cenozoic history the area closed up continent collision zones has resulted in the following late-stage
as the Arabian Plate and Eurasia converged and intervening oceans characteristics of the area undergoing ‘incipient subduction’: 1)
(Neotethys, Sistan oceans) were subducted northwards and lost (e.g. subduction cannot proceed much further from its present state, the
Golonka, 2004). In the south Caspian area the main shortening and remnant of oceanic crust is surrounded by continental crust that is
basin inversion stage occurred during the Late Eocene–Early shortening in collision zones and will prevent large-scale subduction,
Oligocene as the NeoTethys oceanic domains were closed (Kopf although underthrusting will continue, 2) The swamping of the basin
et al., 2003). The Late Miocene–recent phase of compression, uplift by continent-derived sediments deposited in a fluvio-deltaic setting,
and magmatism coincides with the final stages of Arabia –Eurasia and 3) the lateral transition from deepwater folds to shallow or
collision (Nikishin et al., 2001). Mountains adjacent to the South subaerial folds, similar to the Indo-Burma Ranges.
Caspian Sea have been uplifting and providing a sediment source According to Devlin et al. (1999) and Brunet et al. (2003) the South
since the Paleogene. However, rapid Plio-Pleistocene uplift, which Caspian Basin is thought to comprise up to 5 km of Mesozoic
occurred in the Lesser and Greater Caucasus Mountains, the Kopet sediments and volcanics that overlie crystalline basement. The
Dagh Mountains and Alborz Mountains triggered a very high influx Mesozoic units are overlain by marine Paleocene and Eocene marls,
of sediment into the South Caspian Sea (Jackson et al., 2002; Brunet clays, bituminous shales and sandstones of up to 1700 m thickness. A
et al., 2003). tectonic re-arrangement in the basin configuration occurred during
Allen et al. (2002) investigated subsidence in the South Caspian the early Oligocene, in which some basins were eliminated, and
Basin. They noted that the ~2.4 km of tectonic subsidence since 5.5 Ma deposition in the Caspian Sea area occurred in a low-relief setting. The
calculated from backstripping was at a very rapid rate in comparision Oligocene–Early Miocene clay-rich Maykop Series is over 3 km thick
with typical foreland basins, and more than half the sediment and is an excellent hydrocarbon source rock (Katz et al., 2000). The
thickness of the basin was accumulated during the Pliocene– Maykop Series are overlain by the up to 10 km of the late Miocene–
Holocene. To explain such anomalous subsidence, Allen et al. (2002) Holocene Productive Series. Rapid deposition resulted in widespread,
concluded that “subduction” of the South Caspian basement (either high magnitude overpressures, and the extensive development of
thick oceanic crust or thinned, high-velocity continental crust) began mud pipes, diapirs and volcanoes that are mostly sourced from the
ca. 5.5 Ma, as a result of regional reorganization of deformation in the Maykop Series.
Arabian–Eurasian collision zone. Present day convergence rates to the Type 1 near field stress-driven, mobile shale-floored growth faults,
NW are 8–10 mm/yr (Jackson et al., 2002). Knapp et al. (2004) have shale diapirs and fold and thrust belts are present in the eastern part
shown evidence for an oceanic slab extending over 100 km northward of the South Caspian (Turkmenistan), and at least, in part, developed
beneath the Eurasian continental crust (schematically shown in earlier than the Pliocene–Holocene folds that dominate the area
78 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

(Devlin et al., 1999; Lawrence and Babaev, 2000). The spectacular characterize the different categories of DWFTB. The most widespread
folds of the South Caspian Sea vary considerably in map view exploration and development activities have been on salt-associated
geometry, ranging from long (50–150 km) linear, to strongly curved passive margin plays in the South Atlantic (e.g. Gulf of Mexico, Brazil,
(NW–SE to NE–SW strike directions), and dome and basin forms. Angola, Congo, Gabon, Mauritania). There is considerable detailed
Folds in the South Caspian Sea display a similar structural style published information for these plays (see Weimer et al., 2006 for a
whether they are in deep or shallow waters. Unlike most DWFTBs, review), which contrasts strongly with the paucity of data for other
where there is a distinct relationship between water depth and the plays. Consequently the comparisons made below are based on
location of folds, some of the larger fold trends on the western side of the uneven data quality.
South Caspian Basin, which lie predominantly in Azerbaijan, have a
strike length of 150 km, trend NW–SE and pass from being present on 5.1.1. Source rock types
land in the NW to deepwater conditions at the SE part of the trend Large deltas can be up to 10 km thick beneath the shelf, hence
(Fig. 38). Despite the unusual fold-water depth relationship the associated source rocks generally have to be present within the deltaic
province has much in common with DWFTBs because it is affected by sediments. Even if potential source rocks are present in pre-deltaic
shallow gravity driven deformation associated with deltaic deposition, sediments, the thickness of the delta wedge, and generally late
displays growth faults and down-dip folds and thrusts, and has development of the DWFTB precludes section below the basal mobile
extensive, overpressured mobile shales (Devlin et al., 1999; Lawrence salt or shale detachment rocks from being mature at the time of
and Babaev, 2000). Three related factors contribute to the shallow into structural development, except for the outer part of the wedge. For
deep water lateral transition along the fault axes: 1) rapid sedimenta- example in general the Paleogene Agbada–Akata Formations are
tion rates that caused marked infilling of the basin during the Pliocene thought to be the main source rocks for the Niger Delta including the
from fluvio-deltaic systems arranged radially around the basin, 2) the deepwater (e.g. Evamy et al., 1978; Bustin, 1988). Yet the observation
inland sea setting where a remnant of oceanic crust has been preserved from the Gulf of Mexico that the deepwater area is sourced by older
surrounded by continental crust, and 3) the presence of active mountain source rocks than on the shelf led Haack et al. (2000) to propose
belts (Alborz, Kopet Dagh and Caucus mountains) that both act as hypothetically that the deepwater plays of the Niger Delta might be
extensive sediment sources and are associated with Cenozoic regional sourced in part by Late Cretaceous source rocks. Recent geochemical
folding, thrusting and uplift adjacent to, and within the basin. analysis of deepwater Niger Delta oils by Samuel et al. (2009) suggests
The Mesozoic–Eocene section appears to be deformed by thrust that this prediction is probably correct.
faults and possibly strike-slip or oblique slip faults (schematically Source rocks in the deepwater Gulf of Mexico source rocks are
illustrated in Fig. 39). The basement-involved deformation is related primarily located within the Tithonian sequence, but there are
to right-lateral oblique thrusting and inversion of the Caucasus secondary source rocks of Oxfordian and Mid-Cretaceous age (Cole
Trough, which contains very thick sedimentary sequences. Folding in et al., 2001). Additionally, the source rocks in the western Gulf are
the South Caspian Basin displays such a variety of orientations and from carbonate-marl sequences, while source rocks are predomi-
strong vertical disharmony for several key reasons including: 1) nantly carbonate-marl to clastic sequences in the eastern Gulf (Cole et
lateral variations in sediment thickness, 2) tectonic position, 3) al., 2001). These variations in source rock type are reflected in the oil
basement type and structure, and 4) location of overpressured shales, chemistry: the deepwater oils are moderately sour (0.5–20 wt.%
particularly the Maykop Series. sulfur) and predominantly 25–35 API, while ultra-deepwater oils and
Fowler et al. (2000) provide images of a fold in water depths of shelf oils are sweet, low sulfur and high gravity. In contrast to the pre-
300–600 m, across the Shah Deniz structure (70 km SE of Baku). The deltaic source rocks of the Gulf of Mexico, the source rocks for
fold is larger than the other deepwater folds described in this paper. It deepwater hydrocarbons associated with the Baram and Mahakam
is a broad, upright fold in a sedimentary section at least 12 km thick deltas of Borneo appear to have originated from within the deltaic
and, consequently, has a large wavelength of 20 km. A few low- depocentre as a result of re-working of Neogene shallow marine-
displacement thrusts are present at depth. Although the overall estuarine organic material (resins, waxy cuticulae of tropical
structure is simple, in detail the presence of extensive mud volcanoes vegetation, mangrove coals) during lowstands into the deep marine
in the upper 3–5 km of the fold create considerable local complexity. environment (e.g. Saller et al., 2006; Warren et al., in press).
Large mud pipes (3–5 km diameter) are present, and they feed For smaller deltas, and many salt detachment-dominated passive
smaller mud pipes (100's m diameter) and are linked with complex margins, the potential age-range of source rocks is wide since the
associated faulting (such as ring faults, Stewart and Davies, 2006). effects of burial discussed above are not as strong. The South Atlantic
Furthermore, thrusts, normal faults and strike-slip faults located at margin is a very well documented example, where source rocks are
depth make deformation in the lower part of the fold more complex. closely related to the different stages of passive margin development,
The Shah Deniz field is BP's largest field worldwide, it covers a with early continental rift-related lacustrine source rocks, early-drift
surface area of 250 km2, ranges in water depth from 50 m to 600 m, and stage, evaporite associated source rocks, drift-stage Cretaceous
has recoverable reserves estimated at 22 TCF gas and 750 MMBO marine source rocks, and more gas-prone, source rocks deposited by
condensate (BP in Azerbaijan, Sustainability Report, 2005, www.bp. Tertiary deltas (Katz and Mello, 2000; Schiefelbein et al., 2000). The
com/caspian). Large reserves have been also been established in the passive margin setting also means that some source rock intervals
vicinity of the Apsheron Ridge area, most notably the Azeri–Gunashli– may thicken passing offshore. For example Cameron et al. (1998)
Chirag field (432.4 km2, 5–6 billion barrels recoverable; Abrams and attributed a broader suit of Cretaceous and Tertiary oil-prone, algal-
Narmanov, 1997). Yet there remain considerable risks exploring the rich source rocks to expansion offshore of the Atlantic Hinge offshore
large anticlines, and a number have proven to be expensive failures. Angola. The main liability in the deepwater area is that in places the
Coping with overpressure distribution is one problem, reservoir sedimentary overburden of potential source rocks is not sufficiently
distribution/quality and top-seal breach by mud pipes are other factors. thick for their maturation. In some areas of salt tectonics (e.g. West
African margin) hydrocarbon kitchen can be restricted to individual
5. Synthesis of DWFTB characteristics mini-basins between folds or diapirs, causing such areas to have
laterally highly variable source potential. This variation has significant
5.1. Petroleum systems consequences for the hydrocarbon charge to adjacent structures.
The crust underlying the DWFTB in collision zones tends to be
The main aim of this section is to identify broad similarities and thinned continental crust that underthrusts the DWFTB and is buried
differences in the main elements of the petroleum systems that during deformation. Consequently, mature source rocks may be present
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 79

from within the thrust belt wedge, and also below (particularly from flow ~59.0 ± 22.6 mW/m2), but the variations in heat flow value are
syn-rift, post-rift passive margin and foreland basin sequences, e.g. highly complex (Zielinski et al., 2007). These authors suggested that
Ziegler and Roure, 1999). In the collision zones of the Bandar Arc area of high standard deviation in heat flow values reflect hydrothermal
eastern Indonesia, maturation of Paleogene syn-rift sediments is thought circulation, which is similar to the heat and fluid flow regimes of
to be important in explaining the occurrence of oil seeps (Charlton, accretionary complexes. They also note that the heat flow values for
2009). In the Timor Sea, the NW Australian passive margin series with the Baram Delta Province are significantly higher than the generally
their Mesozoic petroleum systems are presently overridden by the low (~ 40–60 mW/m2) mean heat flows associated with accretionary
Timor DWFTB (e.g. O'Brien et al., 1999; Charlton, 2002). According to prisms (e.g. Ferguson et al., 1993). Fluid flow along thrust faults
Baillie (2009), many deepwater areas of Indonesia that were once systems and particularly the basal detachment is a very important
thought non-prospective (e.g. Bone Bay, Cendarawasih Bay, Sumatran mechanism in explaining heat flow, and hydrocarbon migration in
forearc), are now being re-examined since new seismic, multibeam and both convergent DWFTBs and accretionary prism systems (e.g. Fisher
core data has identified extensive areas with thermogenic hydrocarbon and Hounslow, 1990; Zielinski et al., 2007; Barnes et al., 2010). In
seeps that in places can be linked with significant thicknesses of accretionary prisms, heat flow values can increase considerably as the
sedimentary section. Many of these areas are associated with DWFTBs. wedge thins due to fluid flow and expulsion along the basal
In accretionary prisms, the subducting lower plate consists of detachment (Ashi and Taira, 1993).
oceanic crust and lithospheric mantle. Hence, the main potential for
source rocks resides in the upper plate within the thickened 5.1.3. Hydrocarbon migration
accretionary prism (e.g. Makran: Khan et al., 2007; Indo-Burma Ranges: In deepwater areas, the pattern of hydrocarbon migration can be
Maurin and Rangin, 2009; Sumatra: Baillie, 2009). For example, prism assessed by a variety of methods. Seismic reflection data permits
three distinct groups of oils are recognized in the Barbados accretionary mapping of direct indicators of fluid migration such as mud pipes, gas
that are related to generation at multiple levels within the prism from chimneys and various other fluid escape features. These features
Late Cretaceous marine source rocks deposited under dysoxic condi- create sub-vertical zones of low-reflectivity on seismic sections, pock
tions (Hill and Schenk, 2005). Source rocks of this age are widespread marks on time slices, affect seismic attributes (direct hydrocarbon
along the southern Caribbean area west of the trench. indicators, such as bright spots and flat spots) and cause the
preferential accumulation of gas hydrates (marked by bottom
5.1.2. Temperatures and hydrocarbon maturation simulating reflections) in the crests of anticlines; (e.g. Fowler et al.,
Several decades of exploration have established that the conditions 2000; McConnell and Kendall, 2003; Deville et al., 2003, 2006; Gay et
for hydrocarbon maturation in deepwater environments is wide- al., 2006; Stewart and Davies, 2006; Morley, 2009a; Barnes et al.,
spread, and in general, there is nothing specific about the deepwater 2010). Large pipes, 100's m wide, appear to originate deep (kms) in
environment that precludes source rock maturation. However, lower the seismic data, while smaller pipes 10's m across can arise from
geothermal gradients may require a different, generally older, source depths in the order of 100's m (e.g. Morley, 2009a). Overpressured
rock in the deepwater compared with the shelf. Furthermore the fluids may travel through permeable (silty or sandy) beds deep in the
uncertainty associated with source rocks reaching maturation section before rising vertically (e.g. Gay et al., 2006). These permeable
increases as exploration extends into ultra deep areas. beds would filter mud from the fluids leaving them carrying little clay
Published temperature data for DWFTBs are scarce, but some fraction (although they may subsequently incorporate more mud),
generalizations can be made. There is generally a decrease in heat flow unlike the pipes sourcing mud volcanoes. Side scan sonar or multi-
from continental areas to oceanic crust as radiogenic material within beam surveys coupled with targeted piston coring programs, or
the crust is reduced, such as the Amorican Margin, where typical simple grid piston coring programs have also provided important
continental values of heat flow (N90 mW/m2) are higher than oceanic regional coverage that commonly show preferential hydrocarbon
values (b60 mW/m2) (Louden and Mareschal, 1996). Geothermal seepage along specific types of structure, such as young faults, mud
gradients in passive margins reflect more than just the radiogenic heat pipes, salt diapirs and gas chimneys (e.g. Kaluza and Doyle, 1996; Cole
production. The presence of major faults, thick salt and young et al., 2001; Dolan et al., 2004; Zielinski et al., 2007; McConnell et al.,
sediment depocentres can cause significant lateral variations (e.g. 2008; Warren et al., in press).
Forrest et al., 2005; Nagihara and Opre Jones, 2005). Nevertheless There are clear differences between the different types of DWFTBs
diminishing heat flow or geothermal gradients passing offshore are with respect to migration and trap timing, in particular between salt
the general rule. Maturation can be a significant risk for the outer and shale detachment margins, and between margins with landward
parts of DWFTBs on passive margins near the continent–ocean and oceanward-dipping detachments. Types 2a and 2b DWFTB with
transition, as the dual decrease in heat flow and sediment thickness shale detachments are also all associated with a landward-dipping
reduce the source rock maturation potential. However, in some areas detachment and a critical-taper wedge. As described above in the
this change can be advantageous, for example the hotter shelf area in discussion of heat flow, fluid flow within the critical-taper wedges
the Gulf of Mexico is more gas-prone, while the cooler deepwater area tends to be along the thrust faults. Consequently, models for the
is more oil-prone. migration of hydrocarbons in convergent DWFTBs and accretionary
The Niger Delta slope has an average heat flow of 58 mW/m2, with prisms closely resemble each other (Figs. 8B and 40B; Deville et al.,
little decrease in heat flow between the upper and lower slope 2003; Morley, 2009a; Cobbold et al., 2009; Warren et al., in press;
(Brooks et al., 1999). Nevertheless the outer fold and thrust belt is Barnes et al., 2010). Hydrocarbons and overpressured fluids are
characterized by four-way dip-closed anticlines that are either under generated within the wedge, and when the overpressure reaches a
filled by thermogenic hydrocarbons or filled by biogenic gas, critical magnitude fluids move to areas of lower pressure along zones
suggesting that in the outer-most structures, the volume of thermo- of weakness. The wedge geometry forces the fluids to migrate
genically generated hydrocarbons is relatively low (Kostenko et al., oceanward, and the most pervasive zones of weakness tend to be
2008). thrust faults. Hence fluids tend to migrate laterally and upwards along
Convergent setting DWFTBs are underlain by thinned continental thrust fault surfaces until they reach a critical depth where hydraulic
crust. Hence, any offshore increase in depth to the oil and gas fracturing permits a vertical column of overpressured fluid to force its
windows is likely to be due to the increased water depth rather than way to the surface (Fig. 8). Typically fluid pipes develop in the crests of
to a decrease in crustal radiogenic heat production. Heat flow data for anticlines and create mud volcanoes at the sea floor. Fluid pipes and
the Brunei deepwater area suggests mean heat flow decreases down migration of gas up carrier beds along the backlimb of anticlines
slope (upper slope heat flow ~84.0 ± 66.5 mW/m2; lower slope heat appears to focus gas hydrates (as seen by well-defined bottom
80 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

A Growth fault Deepwater fold and thrust belt migration patterns on such margins can be complex, and will have
Depocentre changed considerably spatially and temporally.
Weak migration anticline
In type 1a DWFTBs associated with relatively small deltas,
likely to be underfilled
prograding sedimentary prisms are commonly not thick enough for
the generation of hydrocarbons within them. Consequently, either the
DWFTB needs to be relatively old, contain source rocks, and be buried
to depths suitable for hydrocarbon generation, or a sub-DWFTB source
is required, with vertical migration pathways to the traps. The timing
of sub-DWFTB source rocks is critical. Post-kinematic burial of the
DWFTB may be necessary in order to mature the source rocks within
or beneath them. Alternatively, structural thickening in the DWFTB
B maybe sufficient to bury deeper source rocks to maturity in the largest
deltas. Establishing a vertical migration pathway can be problematic,
and acts contrary to the dip of the basal detachment and dip of the
margin (Fig. 40A). Landward-dipping detachments offer greater
opportunity for the hydrocarbon charge of a DWTFB from source
rocks located both above and below the detachment (Fig. 40B). For
example, the Mexican Ridges play in the Laguna Madre–Tuxpan area
(Neogene folds and reservoirs) requires an underlying Upper Jurassic
Immature source rocks Migration pathway source (Ambrose et al., 2005). Deep-seated basement faults provide
Mature source rocks vertical migration pathways in the southern part of the Laguna
Madre–Tuxpan area, but further north, the offshore-dipping low-
Onset of oil window
angle detachment system above the source rock is likely to form an
Fig. 40. Model illustrating the effect of the dip of the basal detachment surface on effective barrier to hydrocarbon migration into the DWFTB (Ambrose
hydrocarbon migration pathways. A) Seaward dipping detachment, B) Landward et al., 2005).
dipping detachment.

5.1.4. Reservoir quality


The first-order concern pertaining to reservoir quality in DWFTBs
simulating reflections) in the crests of anticlines. The fluid migration is related to the sediment source, which in turn is related to the
events can be pulsed, and the geochemistry of oils from fields in the tectonic setting (Dickinson and Suczek, 1979). In general, the main
Niger Delta and seep studies in Barbados shows variations that are DWFTB reservoirs are turbidite sandstones reworked from clastic
consistent with differing degrees of maturity or biodegradation shelf deposits that were transported to the deepwater areas during
associated with multiple charging events (e.g. Katz, 2003; Hill and lowstand in sea level events or tectonic uplift events (Ofurhie et al.,
Schenk, 2005). 2002; Fainstein, 2003; Ingram et al., 2004; Saller et al., 2008). A large
There is reason for concern that insufficient charge may get to the number of Type 1 DWFTBs are associated with passive margins, such
most distal folds in DWFTBs like Brunei and Niger (Morgan, 2003, as those of eastern North and South America and Africa. These
Kostenko et al., 2008). The seep map for offshore Brunei shows only margins are fed by large river systems, which drain the continental
weak oil seeps in the lower slope area (Fig. 9). A detailed investigation interior. While the geology of the continental interiors is complex and
of the causes of limited column heights in the outer DWFTB of the highly variable, generally a considerable component of the sediments
Niger Delta concluded that insufficient hydrocarbon charge, and not are derived from acidic basement, or Paleozoic–Cenozoic clastics and
trap integrity, was the main problem (Kostenko et al., 2008). This carbonates. The long fluvial transport paths tend to eliminate the
conclusion suggests lateral migration can be of limited efficiency and/ more reactive (generally more basic-ultrabasic igneous or metamor-
or that the source rocks are too lean to generate sufficient phic) minerals, and consequently in general passive margin deepwa-
hydrocarbons to fill the available trap capacity. ter turbidites are predominantly composed of high reservoir quality
Passive margin salt provinces are characterized by a very different quartzose, or quartzo-feldspathic sands, characterized by high
migration scenario in comparison with the vertical to ocean-wards porosities and permeability (e.g. Dickinson and Suczek, 1979;
migration dominated critical taper wedge settings described above. Marsaglia et al., 1996; Fontanelli et al., 2009).
The overall oceanwards dip of the margin means hydrocarbon Type 2 DWFTBs are associated with tectonically active regions
migration is generally either vertical or landwards. As salt is an where the contribution of more highly diagenetically reactive
effective migration barrier, source rocks must either overlie the salt minerals to the sand composition can be much higher, and quartz
unit, or depositional gaps in the salt, large faults, or areas of total salt content is commonly less than 50% (e.g. Dickinson and Suczek, 1979;
withdrawal must be present to permit vertical migration of hydro- Dickinson, 1982). The uplifted hinterlands of many, if not all, the Type
carbons generated by sub-salt source rocks into post-salt structural 2 DWFTB provinces, both convergent zones and accretionary prisms,
traps (e.g. Jackson et al., 1994). Cretaceous and Cenozoic post-salt include associated ophiolites and basic-intermediate volcanic pro-
source rocks underlie effective petroleum systems in large parts of the vinces. Consequently, the likelihood of good quality reservoir
South Atlantic margins (Cameron and White, 1999). However, the development is generally low. The poor quality, volcano-clastic
charge of post-salt prospects by hydrocarbons generated by pre-salt Neogene sandstones of Western Sulawesi (Bergman et al., 1996) are
syn-rift source rocks is important in many areas. Regions of thick salt, the highest risk factor concerning exploration in the convergent zone
such as the salt plateau area of the Outer Kwanza Basin (Fig. 18), have DWFTB. Conversely, the deltas of the NW and eastern Borneo margin
major problems with hydrocarbon charge since migration from sub- have high quality, quartz-rich deepwater sandstones (e.g. Ingram et
salt sources is blocked. The post-salt sediments are generally too thin al., 2004; Saller et al., 2008). As discussed in Section 4.2.1.1, the reason
to be mature, and lateral migration pathways would have to be very for this high reservoir quality is that the older accretionary prism
tortuous. The prolonged history of movement of salt on passive sediments were uplifted to form the main sediment source area, and
margins (over 100 My in the case of West Africa), the ability of salt to were themselves predominantly supplied by axially transported
migrate to different stratigraphic levels with time, and the changing sands (axially along the trench) from an uplifted continental area to
thickness of sediment overburden with time, mean that hydrocarbon the west and southwest. Volcanic and lithic-rich sediments were
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 81

transported westwards from a separate source area into the 2000; Hudec and Jackson, 2004). The transition to deepwater
accretionary prism, but were considerably diluted by sediments deposition can occur over tens of millions of years permitting
derived directly or indirectly from the continental source area. reservoirs to develop in a range of environments, including shelfal
The hypothesis that, under some circumstances, accretionary carbonates as well as clastics. Consequently the range of reservoir
prisms can involve high-quality reservoir rocks is discussed in the types is much greater in salt DWFTBs than other settings. In most
introduction to Section 4.3 Type 2bii Accretionary Prisms. The shale-detachment deepwater folds the reservoir sands were deposited
justification for describing the Barbados, Makran and Indo-Burma prior to folding. Syn-kinematic turbidites tend to onlap the fold crests,
accretionary prisms is their potential for economically viable reservoir and occur in shallow piggyback or mini-basins only, and hence they
rocks. In these areas, major fluvio-deltaic systems, with source areas are not penetrated by crestal wells (e.g. Ingram et al., 2004; Morley and
in continent–continent collision belts, have supplied the respective Leong, 2008; Morley, 2009a). There are a greater variety of trap types
accretionary prism with sediments and thus offer an alternative to the in salt related DWFTBs, for example syn-kinematic turbidite sands can
typical accretionary prism sedimentary source terranes. contain economic hydrocarbon accumulations in mini-basins in up-
Even when good quality reservoir is present, turbidite sandstone dip pinch-out traps (e.g. Flemings and Lupa, 2004).
reservoirs are commonly strongly compartmentalized both by A final feature of salt-related DWFTBs is that fold growth begins
structural features (faults, and possibly shale intrusions; Morley, early in the post-salt depositional history, is slow, and the fold
2003b), diagenesis, and sand body geometry (particularly for wavelength tends to grow with time (Rowan et al., 2004). Since the
channelised, and channel–levee complex sandstones, e.g. Weimer et folds form positive features at the sea floor the growth of salt related
al., 2006; Saller et al., 2008). A number of deepwater projects (perhaps folds can affect sand distribution and transport pathways over a much
most notably Thunder Horse and Mad Dog in the Gulf of Mexico) have longer time-period (sometimes episodically over a 100 my time span)
not produced at the peak rates anticipated, and production has than shale-detachment DWFTBs (generally up to ~ 10 my).
declined much more rapidly than predicted, both as a consequence of
the reservoir complexity, strength of the water drive, and technology 5.1.5. Structural traps
problems arising from remote, high pressure, high temperature There are more similarities than differences in structural style
conditions (Cohen, 2007; http://www.theoildrum.com/node/6415). between the different types of DWFTB, except for the salt detach-
Given the enormous investment required, the unanticipated problems ments. The main characteristics have been listed in Table 3 and are
with understanding even the largest deepwater reservoirs, presents a based on the descriptions of the DWFTBs in the previous sections. In
most significant challenge to the industry. DWFTBs, four-way dip-closed anticlines or three-way dip closures
It was noted previously that the long, slow development of salt- with a lateral thrust seal are the most likely structures to contain
associated DWFTBs impacted the development of migration pathways. economically viable hydrocarbons. Other types of prospective closure
This slow development also influences the reservoir associations of can include oblique or lateral fault seals, and footwall prospects.
salt-related DWFTBs compared with all types of shale-related DWFTBs. Four-way dip-closed anticlines typically arise from detachment
Salt on passive or rifted margins was deposited in a restricted marine folds with no associated thrusts in either limb (Figs. 22, 23),
environment, consequently the overlying sediments tend to range detachment folds with break thrusts, or fault propagation folds
from marginal to shallow marine to deepwater depositional environ- (Fig. 8B; Fig. 41). Salt diapirism may significantly modify the basic fold
ments (e.g. Guardado et al., 1989; Spathopoulos, 1996; Katz and Mello, geometry (Fig. 19). While detachment folds occur above both salt and

Table 3
Summary of key characteristics of DWFTBs. Fold and thrust style, I = imbricate thrusts, FPF = fault propagation fold, FBF = fault bend fold, DF (TF) = detachment fold associated with
thrust fault, PU = pop-up, BT = break thrusts (associated with folding), CF = crestal normal faults. Overpressure, BDC = burial disequilbrium compaction, C = chemical,
I = inflationary, S = slab-derived fluids.

Type 1a Type 1b Type 1 Type 2a Type 2bi Type 2bii

Shale detachment Salt detachment Very large deltas

Stress Near Field Near Field Near Field Near and far field Predominantly Predominantly far field
far field
Presence of large delta Yes Generally no Yes Yes No Occasionally
Detached style Yes Yes Yes Yes Yes Yes
Detachment dip direction Oceanward Oceanward Landward Landward Landward Landward
Basement-involved deformation No No No No known example, Yes No (?)
below DWFTB though it is possible
Mobile unit deformation style Absent or simple Salt nappes, Salt nappes, canopies, Mostly mud pipes, Mostly mud Mostly mud pipes, but
shale diapirs, canopies, diapirs, diapirs, squeezed complex, occasional pipes stratified accretionary
mud pipes squeezed anticlines anticlines, complex and large mobile shale complexes may pass
simple shale diapirs masses (Sabah) laterally into zones of
chaotic deformation (1)
Inversion of DWFTB-related No Yes, in deepwater Uncertain, Deepwater no, No Deepwater no, nearshore yes
growth faults probably no (2) nearshore yes
Propagation of DWFTB Oceanward Landward propagation Oceanward, lateral Oceanward Oceanward Oceanward
can be important (3) migration of DWFTB(4)
Duration 1's my–10's my Episodic up to b 40 my? b 15 my? (5) b15 my? 10's my–100 my+
~ 100my duration
Critical taper 3°b 0–3° 2°–3° 3°–8° 3°–11° 4°–11°
Fold and thrust style I, FPF, FBF, DF(TF) DF, I, FPF, PU, CF DF, I, PU, CF, FBF, FPF DF (TF), I, FPF, FBF DF (TF), I, DF (TF), I, FPF, FBF.
FPF, FBF
Fold wavelength* 1–12 km 2–15 km 2–15 km 5–15 1–30 km 2–30
Dominant vergence Offshore Mixed Offshore Offshore Offshore Offshore
Overpressure BDC, C BDC, C BDC, I , C BDC, I, C BDC, I, C BDC, I, C, S
Shortening amount (km) b20? b 100(7) b 30 (8) b 30 (?) b30? 10's–100's
Other features Inflation of basal
salt layer (9)
82 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

A B
Simple-shear fault bend fold Pure-shear fault bend fold

C D
Simple-shear
Break thrust Classic fault bend fold model
detachment fold

E F
Footwall syncline Thickening by a range of With time a more discrete
small-scale structural processes detachment may develop
within core of fold

Fig. 41. Forward models from starting points A) and B), of C) simple-shear fault-bend folds, E) pure-shear fault bend folds, E) simple-shear detachment fold, and F) classic fault bend
folds. Simple-shear folds do not require slip along a discrete detachment; a detachment zone undergoes an externally imposed bedding-parallel simple shear. In pure-shear fault-
bend folds, slip occurs along a discrete detachment, and a zone in the hangingwall undergoes shortening and thickening above the fault ramp. In the detachment-folding example,
heterogeneous shear occurred, with a basal detachment zone (like the simple-shear fault bend fold), but also layer thickening (like the pure-shear model). The formation of a break
thrust after initial folding results in folding and thickening in both the hangingwall and footwall areas of the break thrust. A–D and F based on Suppe et al. (2004), and Corredor et al.
(2005), E) from Morley, 2009b.

shale detachments, they are most frequently found above salt


detachments (Figs. 22 and 23). The Perdido Fold Belt, Gulf of Mexico
stands out as a particularly well-developed example (e.g. Peel et al., 200
1995; Trudgill et al., 1999). In a deepwater setting where risk
mitigation is vital, simple, large four-way dip closures are the most
prospective structures. As discussed by Trudgill et al. (1999) and
Kostenko et al. (2008), thrusts are sometimes over-emphasized in
seismic interpretations, so that many folds once interpreted as being 150
affected by thrusts in their limbs are now often interpreted as
First order fold length (km)

detachment folds without thrusted limbs.


Most deepwater folds above a shale detachment are associated
with thrusts and commonly a component of lateral thrust fault closure
is required to render such prospects viable. The risks associated with
lateral seal are discussed below in the section on seal (Section 5.1.6). 100
Trap size is a fundamental parameter that may systematically vary
between the different types of DWFTB. Hamilton and De Vera (2009) 82 km
presented data for 23 DWFTBs that showed a considerable variation in 73 km
along-strike fold length and fold wavelength between Type 1a 60 km
(passive margin shale detachment), Type 1b (passive margin salt
50
detachment) and Type 2b (active margin shale detachment) DWFTBs. 44 km 44 km
40 km
These authors concluded that the average fold strike-length of salt-
detachment systems is half that of shale-detachment systems, whilst
average fold amplitudes and wavelengths were similar. For detach- 17 km
ment and buckle folds, the wavelength is strongly controlled by the
thickness (and rheological characteristics) of the controlling compe- 0
Indo-Burma Ranges

South Caspian Sea

Seram

NW Borneo

Makassar Straits

Zagros

Perdido

Nile Delta (Messinian)

Gabon (Astrid)

tent stratigraphic member. Hence stratigraphic thickness of the


section overlying the detachment is likely to have the most significant
impact on fold wavelength. Provinces that tend to have shallow
depths to detachment (e.g. Type 1ai and 1aii DWFTB) will generally
have shorter wavelengths than provinces prone to have deeper
depths to detachment (e.g. Type 1b and 2). The examples of the
wavelength range for large folds, as given in Fig. 11, indicate that the
structural setting and detachment lithology are largely immaterial. Fig. 42. First-order fold length in DWFTBs. Solid lines are lengths of folds with one main
The two largest wavelength provinces (South Caspian Sea, Zagros crest. Dashed lines are folds with multiple crests assumed to be formed by along-strike
Mountains) are both associated with 10 km + thickness of sedimen- linkage.
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 83

tary rock overlying the detachment. Despite not being a DWFTB, the cause the development of lateral and oblique faults (e.g. Niger delta
Zagros Mountains are shown in Fig. 42 since they are the best Cobbold et al., 2009; Morley, 2009b).
representative of a thick sedimentary sequence developed above a In deepwater environments, shale top seals are capable of
salt detachment. However, they also contain thick, Mesozoic and holding back large hydrocarbon columns (Kostenko et al., 2008).
Cenozoic carbonates that form a competent beam, which differentiate However, the top seal can be breached by fluid plumes and mud
them from most DWFTBs. volcano feeder pipes, which are commonly found at the crests of
Figure 42 shows the length of anticlines for a number of thrust and anticlines (e.g. Deville et al., 2006; Morley, 2009a; Warren et al., in
fold belts. The dashed line shows the length of the longest anticlines in press, Fig. 8B). Crestal normal faults resulting from tangential
the respective belt. Such anticlines are generally characterized by 3–5 stresses or from gravity-related stresses (e.g. Morley, 2007b) can
axial culminations and represent composite structures. The solid black also breach top seals.
lines represent the length range for a single periclinal fold or fold According to Ingram et al. (2004), discoveries in the deepwater of
segment. Hamilton and De Vera (2009) noted that salt DWFTBs (Type Sabah demonstrate that sufficient recent oil and gas charge was
1b) were associated with relatively short anticlines, yet the great available to fill young structures. These discoveries were made in
length of the Zagros anticlines indicates that there is nothing intrinsic inactive, buried structures, the trap and seal integrity of which was
to salt detachments that causes short fold lengths. While fold apparently maintained. In the same area there are other, more
wavelength and along-strike length are related to fold growth, other actively growing folds that have not yet been tested, which may have
factors can affect the lateral continuity of folds. These include amongst a greater trap integrity risk due to high fluid pressures and top seal
others oblique and lateral ramps in the basal thrust, relief at the hydrofracturing (Ingram et al., 2004). Similarly the Bobo discovery in
detachment level due to underlying structures (e.g. syn-rift fault the Niger Delta is on an inactive, buried anticline, while an adjacent,
systems) salt diapirs and welds, and lateral discontinuities in the basal recently active structure shows fluid venting to the surface (Cobbold
detachment horizon (salt, or shales). The low fold length in Type 1b et al., 2009).
DWFTBs suggests that in passive margins salt detachment has to
contend with more along-strike obstacles than shale detachments. 5.2. Structural development in different tectonic settings
These obstacles are primarily related to the structural setting under
which the salt-bearing sequence was deposited (early post-rift, or The different tectonic settings described in this paper might be
syn-rift). expected to produce very different DWFTBs. However, broad
While the majority of compressional structures above shale comparisons suggest more similarities than differences (Table 3),
detachments are folds and related thrusts, simple fold or fold and particularly between DWFTBs with a shale detachment irrespective of
thrust belts above a salt detachment are more unusual. As the Gulf of their tectonic framework. The biggest differences occur not between
Mexico demonstrates, much more shortening in the outer wedge is tectonic settings, but between salt and shale detachments. The key
accommodated by diapirism, salt nappe and canopy emplacement differences include:
than by thrusting and folding. In this respect large classic salt
detachment fold-thrust provinces such as the Zagros do not resemble 1) Deformation above salt detachments occurs as a result of
DWFTBs. continental margin post-rift subsidence and loading by sediments
(deltas in particular) combined with uplift of the hinterland.
5.1.6. Seal Deformation of a passive margin sedimentary prism above a shale
Perhaps the greatest single risk inherent to deepwater structural detachments results either from far-field stresses or sediment
prospects are uncertainties about the seal integrity. Four factors loading; margin uplift is less significant as a driving mechanism.
influencing seal integrity are considered here: top seal quality, oblique 2) Salt material properties, combined with a thick sediment overbur-
and lateral faults, thrust fault seal and hydraulic fractures. den, permit the development of salt diapirs, salt nappes and salt
In Type 1a DWFTBs, prospective structures are typically associated canopies. These features can interfere with the development of
with shale or mudstone top seals, and if faults contribute towards regular folds and thrusts in DWFTBs. Thrusts and imbricate wedges
closure, lateral seals are provided either by reservoir/clay juxtaposi- can form in response to the activation of a basal shear along the salt
tion or by membrane seals (clay smears). As discussed above, large allochthon (Hudec and Jackson, 2009). Salt nappe development may
thrust faults are commonly linked with a basal shale thrust control the amount of salt available for fold detachments (Perdido vs.
characterized by very high overpressures. These faults provide Mississippi Fan fold belts). Detachment-type fold belts with little
conduits for fluids to rise and form feeder pipes for mud volcanoes associated thrusting are a feature of salt detachments.
(e.g. Deville et al., 2006; Morley, 2009a; Barnes et al., 2010; Fig. 8). 3) Mobile shale provinces produce more extensive tracts of folds and
Hence faults are a major risk as lateral seals. The young development thrusts than salt detachments, and cannot develop extensive
of many fold and thrust belts means that fault reactivation is a nappes and canopies. Their equivalents are probably recycled
significant liability for many structures. Older faults in the DWFTB that mudstones (originating from mud volcanoes) in the syn-kinematic
have been inactive for some time have more potential to act as lateral series. Even large shale cored diapirs may be rare features. With
seals. increasing seismic resolution, most “shale diapirs” are revealed as
Lateral and oblique faults appear to be commonly associated with steep-flanked, tight anticlines that in some cases are pierced by
hydrocarbon seeps, as evidenced by deepwater coring programs. pipe-like features feeding mud volcanoes (Van Rensbergen et al.,
Hence, if such faults form a component of a structural closure, or 1999). Shale detachment provinces are associated with extensive
control the maximum hydrocarbon column height, their reactivation piercement features due to overpressured fluids (mud pipes, mud
risk is likely to be more significant than that of the frontal thrust. For chambers, mud volcanoes, gas chimneys). Changes in deformation
example, oblique faults are associated with the majority of hydrocar- style in shale detachment provinces are, in part, related to the
bon seeps in the Barbados Accretionary Prism (Dolan et al., 2004). The nature of the detachment, whether it forms a thin, discrete zone of
occurrence of oblique and lateral faults is highly variable. In some sliding, or a thick mobile shale zone. Generally, shale detachment
settings they can be closely spaced (e.g. accretionary prisms, Dolan et folds are associated with thrusts (usually fault bend fold, fault
al., 2004; Barnes et al., 2010, Type 1a slides, De Vera et al., 2010, and propagation fold, or break-thrust styles).
thin sedimentary sequences overlying salt detachments, Clark and 4) Vergence of structures above mobile shale detachments tends to
Cartwright, 2009). Local topographic features, such as seamounts and be seaward, although local variations can occur. The vergence of
transform faults, that may lie below the basal detachment can also structures above highly overpressured shale detachments is
84 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

sometimes landwards. The vergence of structures above salt (millions of years) for small deltas, may last for tens of millions of
detachments is considerably more variable. years for large deltas. Upon termination of delta progradation, the
5) In salt detachments the sequence of deformation can be ocean- loss of overpressure along the detachment results in the
ward, but along the African margin buttressing of the salt province termination of DWFTB deformation. However, salt detachments
in the deepwater area resulted in landward propagation of the are able to keep reactivating simply due to material properties and
compressional belt (Hudec and Jackson, 2004; Jackson et al., 2008). hence, the duration of activity (albeit episodic) can exceed 100 My
In a shale detachment setting, deformation tends to propagate even in areas where no major delta provides a continuous
oceanward, although synchronous and out-of-sequence deforma- sediment supply to the margin.
tion within the wedge is common (e.g. Morley, 1988,2003a; Moore
Differences in deformation style between the various tectonic
et al., 2007; Cubas et al., 2008).
settings of shale detachment include the following:
6) Shale detachments form DWFTBs that can be treated as critical
taper wedges. Salt detachments do not conform to a clear wedge 1) DWFTBs in a Type 1 setting (both salt and shale detachments) are
geometry (e.g. Davis and Engelder, 1985; Dahlen, 1990; Bilotti and characterized by a landward (shelfal) zone of extension (margin
Shaw, 2005; Cubas et al., 2008; Fig. 18). sub-parallel SHmax direction) and an oceanward (slope) zone of
7) Overpressures are generally trapped below salt units. The onset of compression (margin sub-perpendicular SHmax direction;
high overpressures is abrupt (sometimes over a thickness of Fig. 43). DWFTBs in a Type 2a setting have an onshore-inner
centimeters) at the base of the salt unit (e.g. Mostofi and Gansser, shelf zone of compression/inversion, a narrow shelfal zone of
1957). In shale detachments overpressures progressively increase active extension and an oceanward (slope) zone of compression,
within shales approaching the detachment. Highest overpressures reflecting mixed near-field and far-field stresses affecting the
may be attained within the detachment, not below it. margin. Types 2aii and 2b DWFTBs are generally characterized by
8) On passive margins, shale detachment activity is related to the compression throughout the wedge, because there is no near-field
location and duration of delta progradation. This can be short effect from deltas.

A B C

I I
I

X Y
II II

II Z III
III

Near-field stress Mixed near- and far- Far-field stress driven


driven systems field stress driven systems systems
IV III IV

Maximum and Maximum and


minimum horizontal minimum horizontal
Extensional fault Thrust
stress orientation stress orientation
(extensional stress) (compressional stress)

Fig. 43. Diagram illustrating idealized cross-sections and stress orientations in three main types of DWFTBs. A = near field stress systems, AI) gravity sliding on oceanward-dipping
system, AII) mixed oceanward and landward-dipping system, AIII) salt-detachment with diapirism and salt nappe formation (salt in black), AIV) map view of stress orientations.
B = mixed near-and far-field stresses, BI) continental thrust belt, (or possibly accretionary prism) linking with near-field stress system towards foreland, BII) collapse of thickened
continental crust driven by flow of crust at base towards hinterland (Z), and ‘thick-skinned’ far-field gravity collapse of crust (Y) linking with ‘thin-skinned’ near-field gravity driven
deformation within the sedimentary section (X), BIII) map view of stress orientations. C) Far-field stress diven systems, CI) accretionary prims, CII) continent–arc, or continent–
ophiolite collision, CIII) intracontinental convergent zone (inversion and thin-skinned thrusting of rift-post-rift basin section on margin of collision zone, CIV) map view of stress
orientations. Grey = sedimentary section involved in DWFTB.
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 85

2) Far-field stresses tend to produce both thick-skinned and thin- highlights that DWFTBs can develop during all the stages of the
skinned deformation; in particular normal fault-bounded basins Wilson Cycle, from the passive margin stage, through the ocean
are prone to inversion in Type 2, but not Type 1 settings. subduction stage to early-stage of continent–continent collision, and
3) The period of DWFTB development in Type 2a and 2bi settings can in remnant post-collision basins such as the Mediterranean and Black
be relatively short (around 10–15 My) since continent collision is Sea. The second subdivision of the proposed classification distin-
acting to shorten and uplift the DWFTB. The life span of individual guishes between salt and shale detachments. However, for Type 2
DWFTBs on Type 1 margins also tends to be less than 10 My either settings, there are no examples of modern DWFTBs associated with
because the activity of small deltas only lasts a few million years, or evaporite detachments. Consequently,Type 2 detachments are only
because older DWFTBs become overstepped and buried during shown with a shale detachment in Figure 2. The sub-division of near-
progradation of large deltas (which develop over tens of millions field stress systems into salt and shale detachments is very important
of years, and the present day deformation belt (such as in the Niger because they mark fundamental differences in structural style and, to
Delta) may only represent the last 10 My of development. a lesser degree, differences in duration of tectonic activity (ability of
Accretionary prisms can contain a record of deformation lasting the detachment zone to reactivate) and propagation direction of
many tens of millions of years. deformation (Table 3). Shale detachments are by no means homog-
enous, and further subdivision could consider on how different shale
Similarities between some of the different provinces are also
detachment type affects the structural style (e.g. thin, well compacted
important.
and overpressured; narrow weak minerals and hydrostatic pressure;
1) It is striking that large growth-fault bounded depocentres can be narrow, undercompacted and overpressured; broad, undercompacted
present in all settings in shale-detachment DWFTBs: e.g. Makran and overpressured).
(Type 2bii), NW Borneo (Type 2a), not just in Type 1 (Niger Delta). The largest hydrocarbon provinces in each setting (Type 1a, Niger
However, the occurrences of large growth fault depocentres in Delta; Type 1b; Gulf of Mexico; Type 2a, NW Borneo; Type 2bii,
Type 2 settings appear to be confined to where large deltas have Caspian Sea) are not entirely representative of the more frequently
interacted with far-field driven systems. Growth faults in occurring members of that category. The Gulf of Mexico with its short
accretionary prisms have been previously explained as a product period of sea floor spreading, and sheltered location away from the
of collapse of over thickened critical wedges (Platt, 1986). Such a main Atlantic passive margin where sediments were funneled in from
mechanism may be one part of the explanation, but in the case of Northern America is an atypical example of passive margins. The
the Makran (Fig. 36), to fill a large growth faults depocentre, with a Niger Delta with its broad zone of mobile shale activity (controver-
high proportion of sand-prone, shallow marine sediments also sially called diapirs), squeezed folds, and thick mobile shales is not
implies an active deltaic system was present. Hence, at least for actually typical for most shale-detachment deltas. In NW Borneo, the
certain stages in its development, the Makran may have developed coincidence of a system of deltas with good-quality reservoir sands,
as a linked near-field and far-field driven system, similar to NW building out onto a convergent margin with a DWFTB driven by mixed
Borneo. near-field and far-field stresses is atypical. In the Caspian Sea, the
2) Some shale detachments in Type 1 and 2 settings are associated Absheron Ridge is associated with a most unusual early-stage
with DWFTB structures that show either mixed or landward accretionary prism that developed during a phase of limited
vergence (e.g. northern Indo-Burma Ranges; external fold and subduction of oceanic crust spanning perhaps as little as 10 My. The
thrust belt, Niger Delta; Cascadia accretionary prism). These volume of continent-derived sediment that was deposited in the
detachments are inferred to be super-weak due to very high South Caspian Sea is highly anomalous for accretionary prisms.
pore fluid pressures (Byrne and Hibbard, 1987; Underwood, 2002: For Type 1 DWFTBs, the volume of accumulated sediments is
Steckler et al., 2008; Cobbold et al., 2009). The occurrence of very crucial for their hydrocarbon potential, particularly in terms of source
similar fold geometries in the Indo-Burma Ranges and the external rock development and whether these have been buried sufficiently to
Niger Delta, despite their different tectonic settings, is striking. attain to maturity to expel hydrocarbons during and after the
formation of structural traps. In this respect the hydrocarbon potential
6. Conclusions of smaller deltas, or passive margins without significant deltas may be
limited. Even in DWFTBs with the greatest exploration success (Gulf
The classification of DWFTBs presented here is, in many aspects, of Mexico, Niger Delta), the distribution of commercial hydrocarbon
similar to those of Rowan et al. (2004), Krueger and Gilbert (2009) accumulations is very uneven, raising questions about source rock
and Hamilton and De Vera (2009). It identifies the differences in development and maturity in deepwater environments. The issue of
deformation styles, duration of activity, and amount of shortening, the source rock maturity goes beyond just sediment thickness and also
megatectonic setting and whether the basal detachment zone is includes the general decrease in heat flow passing oceanwards across
located in weak shales or evaporites. Previous classifications assumed passive margins. Smaller Type 1a systems are also prone to develop
that the tectonic setting could be equated to the driving mechanism of shorter-wavelength folds, or folds in the hangingwalls of imbricate
DWFTBs (i.e. passive margins = near field stress driven systems, thrusts, as compared to large deltas. Such characteristics produce
collisional margins = far-field stress driven systems). This paper has relatively small trap volumes or hydrocarbon column heights, and
shown that such distinctions are not so clear-cut, and consequently, render prospects marginal to sub-economic in deepwater environ-
has selected driving mechanism, not tectonic location as the primary ments. Typical salt detachment passive margins tend to have poorly
subdivision of the proposed classification. Of particular note is that developed DWFTBs; by far the best-developed ones are associated
some DWFTBs in collision belts (e.g. NW Borneo), transpressional with the very large deltaic province of the Gulf of Mexico.
zones (e.g. Columbus Basin, Trinidad) and possibly some accretionary Type 2a and 2bi DWFTBs such as NW Borneo and Timor involve
prisms (e.g. the Makran) are driven by mixed near-field and far-field continental crust entering the subduction zone. Consequently the
stresses, and not just by far-field stresses as assumed in previous geothermal gradient passing offshore is not subject to the same
classifications. oceanward decrease as on passive margins. The presence and
The effect of the megatectonic setting of a DWFTBs is particularly maturity of source rocks are still important issues, but may not be
marked in the domains of deformation duration, cause of deformation as critical as on passive margins. Reservoir presence and quality is a
driving stresses, amount of shortening and dimensions of the DWFTB. major concern for both Type 2a and 2b DWFTBs, and is a major
Therefore, the megatectonic setting of DWFTBs is used as the first liability for some Type 2bi DWFTBs (e.g. West Sulawesi) and most
subdivision criterion of the proposed classification. This review Type 2bii DWFTBs. However, reservoir quality need to be addressed
86 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

on a case-by-case basis, since there are examples, such as NW Borneo, Modern Developments in Structural Interpretation, Validation and Modelling:
Geol. Soc., London, Spe. Publs., 99, pp. 251–263.
Trinidad, South Caspian Sea and the Indo-Burma Ranges where good- Belderson, R.H., Kenyon, N.H., Stride, A.H., Pelton, C.D., 1984. A “braided” distributary
quality, continent-derived sands have entered the systems. Proximity system on the Orinoco deep-sea fan. Mar. Geol. 56, 195–206.
to a large continental area is a prerequisite. Even with good quality Berberian, M., 1983. The southern Caspian – a compressional depression floored by a
trapped, modified oceanic crust. Can. J. Earth Sci. 20, 163–183.
reservoirs, compartmentalization can create significant problems for Bergman, S.C., Coffield, D.Q., Talbot, J.P., Garrard, R.A., 1996. Tertiary tectonic and
economic exploitation of the reserves. As a general characteristic, the magmatic evolution of western Sulawesi and the Makassar Strait, Indonesia:
number and size of structural traps is a highly favorable aspect of Type evidence for a Miocene continent-continent collision. In: Hall, R., Blundell, D. (Eds.),
Tectonic evolution of Southeast Asia: Geol. Soc. Spec. Pub., 106, pp. 391–429.
2 DWFTBs, although in some areas relatively small wavelength folds Berhmann, J.H., 1991. Conditions for hydrofracture and the fluid permeability of
can develop, or the intensity of deformation may be too great. A accretionary wedges. Earth Planet. Sci. Lett. 107, 550–558.
significant risk, particularly in Type 2b settings, is the presence of Bigi, S., Lenci, F., Doglioni, C., Moore, J.C., Carminati, E., Scrocca, D., 2003. Decollement
depth versus accretionary prism dimension in the Apennines and the Barbados.
oblique faults (commonly strike-slip faults) that appear to be non-
Tectonics 22. doi:10.1029/2002TC001410.
sealing, leak fluids and breach potential hydrocarbon traps. Bilotti, F., Shaw, J.H., 2005. Deepwater Niger Delta fold and thrust belt modeled as a
critical-taper wedge: the influence of elevated basal fluid pressure on structural
Acknowledgements styles. AAPG Bull. 89, 1475–1491.
Bird, D.E., Burke, K., Hall, S.A., Casey, J.F., 2005. Gulf of Mexico tectonic history: hotspot
tracks, crustal boundaries, and early salt distribution. AAPG Bull. 89, 311–328.
Reviewers Peter Ziegler and Chris Talbot provided very useful, Breen, N.A., Silver, E.A., Hussong, D.M., 1986. Structural styles of an accretionary wedge
detailed and encouraging reviews that must have been very time south of the island of Sumba, Indonesia, revealed by SeaMARC II side scan sonar.
Geol. Soc. Am. Bull. 97, 1250–1261.
consuming. These reviews helped considerably to improve the Bridge, C., Calon, T.J., Hall, J., Aksu, A.E., 2005. Salt tectonicsin two convergent-margin
manuscript and their contribution is gratefully acknowledged. John basins of the Cyprus Arc, Northeastern Mediterranean. Mar. Geol. 221, 223–259.
Warren is thanked for an earlier review of the manuscript. Professors Briggs, S.E., Davies, R.J., Cartwright, J.A., Morgan, R., 2006. Multiple detachment levels
and their control on fold styles in the compressional domain of the deepwater west
Ahmed Chalouan and André Michard are gratefully acknowledged for Niger Delta. Basin Res. 18, 435–450.
promptly providing comments about the tectonic setting of the Prerif Brooks, J.M., Bryant, W.R., Bernard, B.B., Cameron, N.R., 1999. The nature of gas hydrates
zone, and Tony Barber is thanked for his comments about the Timor on the Nigerian continental slope. Annals of the New York Academy of Sciences
Third International Conference on Gas Hydrates, July 18-22, 1999, Park City, Utah.
Sea. The contributions of Ros King, Richard Hillis, Mark Tingay and http://www.tdi-bi.com/our_publications/gas_hydrates/gas_hydrates.html.
Guillaume Backe to this article form TRaX record 109. Brown, K.M., 1990. Nature and hydrogeological significance of mud diapirs and
diatremes for accretionary systems. J. Geophys. Res. 95, 8969–8982.
Brown, K.M., Bekins, B.A., Clennell, B., Dewhurst, D., Westbrook, G.K., 1994.
References Heterogeneous hydrofracture development and accretionary fault dynamics:
reply. Geology 77, 1053–1054.
Aal, A., El Barkooky, A., Gerrits, M., Meyer, H., Schwander, M., Zaki, H., 2000. Tectonic Brun, J.-P., Merle, O., 1985. Strain patterns in models of spreading-gliding nappes.
evolution of the Eastern Mediterranean Basin and its significance for hydrocarbon Tectonics 4, 705–719.
prospectivity in the ultradeepwater of the Nile Delta. Lead. Edge. 19. doi:10.1190/ Brunet, M.-F., Korataev, M.V., Ershov, A.V., Nikishin, A.M., 2003. The South Caspian
1.1438485. Basin: a review of its evolution from subsidence modeling. Sed. Geol. 156, 119–148.
Abrams, M.A., Narmanov, A.A., 1997. Geochemical evaluation of hydrocarbonsand teir Buffler, R. T., 1983. Structure of the Mexican Ridges fold belt, southwest Gulf of Mexico.
potential sources in the western South Caspian depression, Republic of Azerbaijan. In: A. W. Bally, (editor), Seismic expression of structural styles; a picture and work
Mar. Petrol. Geol. 14, 451–468. atlas; v. 2: AAPG Studies in Geol. 15: 2.3.3-16 to 2.3.3-21.
Acharyya, S.K., 2007. Collisional emplacement history of the Naga-Andaman ophiolites Bustin, R.M., 1988. Sedimentology and characteristics of dispersed organic matter in
and the position of the eastern Indian suture. J. Asian Earth Sci. 29, 229–242. Tertiary Niger Delta: origin of source rocks in a deltaic environment. AAPG Bull. 72,
Ahmad, S.S., 1969. Tertiary geology of part of south Makran Balochistan, west Pakistan. 277–298.
AAPG Bull. 53, 1480–1499. Butler, R.W.H., Paton, D.A., 2010. Evaluating lateral compaction in deepwater fold and
Ajakaiye, D.E., Bally, A.W., 2002. Manual and atlas of structural styles. Niger Delta: AAPG thrust belts: how much are we missing from “nature's sandbox”? GSA Today 20 (3),
Continuing Education Course Notes Series 41. 102 pp. 4–10.
Allen, M.B., Jones, S., Ismail-Zadeh, A., Simmons, M., Anderson, L., 2002. Onset of Byrne, T., Hibbard, J., 1987. Landward vergence in accretionary prisms: the role of the
subduction as the cause of rapid Pliocene-Quaternary subsidence in the South backstop and thermal history. Geology 15, 1163–1167.
Caspian basin. Geology 30, 775–778. Byrne, D.E., Sykes, L.R., Davis, D.M., 1992. Great thrust earthquakes and aseismic slip
Allen, R., Najman, Y., Carter, A., Barfod, D., Bickle, M.J., Chapman, H.J., Garzanti, E., along the plate boundary of the Makran subduction zone. J. Geophys. Res. 97,
Vezzoli, G., Ando, S., Parrish, R.R., 2008. Provenance of the Tertiary sedimentary 449–478.
rocks of the Indo-Burman Ranges, Burma (Myanmar): Burman arc or Himalayan- Calvert, S.J., Hall, R., 2007. Cenozoic evolution of the Lariang and Karama regions, North
derived. J. Geol. Soc. 165, 1045–1057. Makassar Basin, western Sulawesi, Indonesia. Pet. Geosci. 13, 353–368.
Ambrose, W.A., Wawrzyniec, T.F., Fouad, K., Sakurai, S., Jennette, D.C., Brown Jr., L.F., Cameron, N.R., White, K., 1999. Exploration Opportunities in Offshore Deepwater
Guevara, E.H., Dunlap, D.B., Talukdar, S.C., Garcia, M.A., Romano, U.H., Vega, J.A., Africa: IBC Oil and Gas Developments in West Africa, London, 25–26 October, 1999.
Zamora, E.M., Ruiz, H.R., Hernandez, R.C., 2005. Neogene tectonic, stratigraphic and 28 pp.
play framework of the southern Laguna Madre-Tuxpan continental shelf, Gulf of Cameron, N.R., Schiefelbein, C.F., Brooks, J.M., Brandao, M.G.P., 1998. The application of
Mexico. AAPG Bull. 89, 725–751. seabed sampling and organic geochemistry to frontier basin exploration in Angola.
Ashi, J., Taira, A., 1993. Thermal structure of the Nankai accretionary prism as inferred 3rd Annual Forum Worldwide Deepwater Technologies, London, 17–18 February,
from the distribution of Gas Hydrate BSRs. In: Underwood, M.B. (Ed.), Thermal 1998. 8 pp.
Evolution of the Tertiary Shimanto Belt, Southwest Japan: An Example of Ridge- Canals, M., Lastras, G., Urgeles, R., Casamor, J.L., Mienert, J., Cattaneo, A., De Batist, M.,
Trench Interaction: Geol. Soc. Am. Spec. Paper, 273, pp. 137–150. Haflidason, H., Imbo, Y., Laberg, J.S., Locat, J., Long, D., Longva, O., Masson, D.G.,
Audley-Charles, M.G., 1986. Rates of Neogene and Quaternary tectonic movements in Sultan, N., Trincardi, F., Bryn, P., 2004. Slope failure dynamics and impacts from
the southern Banda Arc based on micropalaentology. Geol. Soc. Lond. 143, 161–175. seafloor and shallow sub-sea geophysical data: case studies from the COSTA
Baillie, P., 2009. A new prospectivity study puts old doubts to rest. E&P mag. July. project. Mar. Geol. 213, 9–72.
Balaguru, A., Hall, R., 2009. Evolution and Sedimentation of Sabah, North Borneo, Carter, D.J., Audley-Charles, M.G., Barber, A.J., 1976. Stratigraphical analysis of island
Malaysia. Search and Discovery Article 30084. 15 pp. arc-continental margin collision in eastern Indonesia. J. Geol. Soc. Lond. 132,
Bangs, N.L.B., Gulick, S.P.S., 2005. Physical properties along the developing decollement in 179–198.
the Nankai Trough: inferences from 3-D seismic reflection data inversion and Leg 190 Cartwright, J., Huuse, M., 2005. 3D seismic technology: The geological “Hubble”. Basin
and 196 drilling data. In: Mikada, H., Moore, G.F., Taira, A., Becker, K., Moore, J.C., Klaus, Res. 17, 1–20.
A. (Eds.), Proc. ODP, Sci. Results, 190/196, pp. 1–18. Cavazza, W., Roure, F.M., Spakman, W., Stammpfli, G.M., Ziegler, P.A. (Eds.), 2004. The
Barber, A., Brown, K., 1988. Mud diapirism: the origin of mélanges in accretionary TRANSMED Atlas, The MediterraneanRegion from Crust to Mantle. Springer, Berlin.
complexes? Geol. Today 4, 89–94. 125 pp., compact disc.
Barber, A.J., Tjokrosapoetro, S., Charlton, T.R., 1986. Mud volcanoes, shale diapirs, Chalouan, A., Michard, A., El Kadiri, K., Negro, F., Frizon de Lamotte, D., Sotto, J.I., Saddiqi,
wrench faults and mélange in accretionary complexes, eastern Indonesia. AAPG O., 2008. The Rif Belt. In: Michard, A., Saddiqi, O., Chalouan, A., Frizon de Lamotte, D.
Bull. 70, 1729–1741. (Eds.), Continental Evolution: The Geology of Morocco. Advances in geographic
Barnes, P.M., Lamarche, G., Bialas, J., Henrys, S., Pecher, I., Greinert, J., Mountjoy, J.J., information Science, 116. Springer, Berlin, pp. 203–302.
Pedley, K., Crutchley, G., 2010. Tectonic and geological framework for gas hydrates Charlton, T.R., 2002. The petroleum potential of East Timor. APPEA J. 42, 351–369.
and cold seeps on the Hikurang subduction margin, New Zealand. Mar. Geol. 272, Charlton, T.R., 2004. The petroleum potential of inversion anticlines in the Banda Arc.
26–48. AAPG Bull. 88, 565–585.
Beekman, et al., 1996. Crustal fault reactivation facilitating lithospheric folding/ Charlton, T., 2009. The Banda Arc: Geology and Petroleum Potential. http://www.
buckling in the Central Indian Ocean. In: Buchanan, P.G., Nieuwland, D.A. (Eds.), manson.demon.co.uk/bandaforearcstru.html.
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 87

Cita, M.B., 1983. The Messinian salinity crisis in the Mediterranean: a review. In: Duerto, L., McClay, K., 2002. 3D Geometry and evolution of shale diapirs in the Eastern
Berckhemer, H., Hsu, K.J. (Eds.), Alpine Mediterranean Geodynamics Review: Am. Venezuelan Basin. AAPG Search and Discovery article #10026. http://www.
Geophys. Union, Geodynamics Series, 8, pp. 113–140. searchanddiscovery.net/documents/duerto/index.htm.
Clark, I.R., Cartwright, J.A., 2009. Interactions between submarine channel systems and Duval, B., Cramez, C., Jackson, M.P.A., 1992. Raft tectonics in the Kwanza Basin, Angola.
deformation in deep fold belts: examples from the Levant basin, Eastern Mar. Pet. Geol. 9, 389–404.
Mediterranean sea. Mar. Petrol. Geol. 26, 1465–1482. Eisenstadt, G., De Paor, D.G., 1987. Alternative model of thrust-fault propagation.
Clift, P., 2006. Controls on the erosion of Cenozoic Asia and the flux of clastic sediment Geology 15, 630–633.
onto the ocean. Earth Planet. Sci. Lett. 241, 571–580. Embly, R.W., 1976. New evidence for occurrence of debris flows deposits in the deep
Cloetingh, S., Beekman, F., Ziegler, P.A., Van Wees, J.-D., Sokoutis, D., 2008. Post-rift sea. Geology 4, 371–374.
compressional reactivation potential of passive margins and extensional basins. In: Escalona, A., Mann, P., Bingham, L., 2008. Hydrocarbon exploration plays in the Great
Johnson, H., Dore, A.G., Gatliff, R.W., Holdsworth, R.E., Lundin, E.R., Ritchie, J.D. Caribbean Region and neighbouring provinces. Search and Discovery Article 10147.
(Eds.), The Nature and Origin of Compression on Passive Margins: Geol. Soc., Posted July 2, 2008.
London, Spec. Publ., 306, pp. 27–70. Evamy, B.D., Haremboure, J., Kamerling, P., Knaap, W.A., Molloy, F.A., Rowlands, P.H.,
Cobbold, P.R., Szatmari, P., 1991. Radial gravitational gliding on passive margins. 1978. Hydrocarbon habitat of Tertiary Niger delta. AAPG Bull. 62, 1–39.
Tectonophysics 188, 249–289. Fainstein, R., 2003. Comparative response of seismic signatures in deep-water
Cobbold, P.R., Mourgues, R., Boyd, K., 2004. Mechanism of thin-skinned detachment in reservoirs offshore Brazil and West Africa. AAPG Annual Convention, Salt Lake
the Amazon Fan: assessing the importance of fluid overpressure and hydrocarbon City, Utah, 11–14 May, 2003, Abstract. 6 pp.
generation. Mar. Pet. Geol. 21, 1013–1025. Faugères, J.C., Gonthier, E., Griboulard, R., Massé, L., 1993. Quaternary sandy deposits
Cobbold, P.R., Clarke, B.J., Loseth, H., 2009. Structural consequences of fluid and canyons on the Venezuelan margin and south Barbados accretionary prism.
overpressure and seepage forces in the outer thrust belt of the Niger Delta. Pet. Mar. Geol. 110, 115–142.
Geosci. 15, 3–15. Feng, J., Buffler, R.T., Kominz, M.A., 1994. Laramide orogenic influence on late Mesozoic-
Cohen, D., 2007. The Gulf of Dispair? Energy Bull. June 27th. http://www.energybulle- Cenozoic subsidence history, western deep Gulf of Mexico basin. Geology 22,
tin.net/node/31378. 359–362.
Cole, G.A., Yu, A., Peel, F., Requejo, R., DeVay, J., Brooks, J., Bernard, B., Zumberge, J., Ferdian, F., Hall, R., Watkinson, I., 2010. Structural re-evaluation of the North Banggai-
Brown, S., 2001. Constraining source and charge risk in deepwater areas. World Oil Sula area, Eastern Indonesia. Proceedings of the Petroleum Systems of SE Asia and
Mag. 222, 69–77. Australasian Conference, May, 2010. Indonesia Petroleum Association. IPA10-G-
Colten-Bradley, V.A., 1987. Role of pressure in smectite dehydration: effects on 009, 20 pp.
geopressures and smectite-to-illite transformation. AAPG Bull. 71, 1414–1427. Ferguson, A., McClay, K., 1997. Structural modelling within the Sanga Sanga PSC,
Condie, K.C., 1997. Plate tectonics and crustal evolution4th edition. Butterworth- Kutei Basin, Kalimantan: its application to paleochannel orientation studies and
Heinemann. 282 pp. timing of hydrocarbon entrapment. Proceedings of the Petroleum Systems of SE
Corredor, F., Shaw, J.H., Bilotti, F., 2005. Structural styles in the deep-water fold and Asia and Australasian Conference, May, 1997. Indonesia Petroleum Association,
thrust belts of the Niger Delta. AAPG Bull. 89, 735–780. pp. 727–749.
Cossey, S.P.J., 2004. Deepwater discoveries toasted over 30 years: AAPG Explorer. Ferguson, I.J., Westbrook, G.K., Langseth, M.G., Thomas, G.P., 1993. Heat flow and
http://www.aapg.org/explorer/2004/09sep/gom_history.cfm. thermal models of the Barbados Ridge accretionary complex. J. Geophys. Res. 98,
Cottam, M., Hall, R., Sperber, C., Armstrong, R., 2008. The Mount Kinabalu granite of 4121–4142.
North Borneo: result and cause of orogenic deformation. Am. Geophys. Union, Fall Fisher, A.T., Hounslow, M.W., 1990. Transient fluid flow through the toe of the Barbados
Meeting, 2008. abstract #T23C-2059. accretionary complex: constraints from the Ocean Drilling Program Leg 110 heat
Cottam, M., Hall, R., Sperber, C., Armstrong, 2010. Pulsed emplacement of the Mount flow studies and simple models. J. Geophys. Res. 95, 8845–8858.
Kinabalu granite, northern Borneo. J. Geol. Soc. Lond. 167, 49–60. Fisher, A.T., Zwart, G., 1997. Packer experiments along the decollement of the Barbados
Cubas, N., Leroy, Y.M., Maillot, B., 2008. Prediction of thrusting sequences in accretionary complex: measurements of in situ permeability. In: Shipley, T.H.,
accretionary wedges. J. Geophys. Res. 113. doi:10.1029/2008JB005717. Ogawa, Y., Blum, P., Bahr, J.M. (Eds.), Proc, Ocean Drill. Prog., Sci. Res., 156,
Cullen, A.B., 2010. Transverse segmentation of the Baram-Balabac Basin, NW Borneo: pp. 199–218.
refining the model of Borneo's tectonic evolution. Pet. Geosci. 16, 3–29. Fisher, A., Zwart, G., the Leg 156 Scientific Party, 1996. The relation between
Curiale, J.A., Covington, G.H., Shamsuddin, A.H.M., Morelos, J.A., Shamsuddin, A.K.M., permeability and effective stress along a plate-boundary fault, Barbados accre-
2002. Origin of petroleum in Bangladesh. AAPG Bull. 86, 625–652. tionary complex. Geology 24, 307–310.
Curray, J.R., Moore, D.G., 1971. Growth of the Bengal Deep-Sea Fan and denudation in Flemings, P.B., Lupa, J.A., 2004. Pressure prediction in the Bullwinkle Basin through
the Himalayas. Geol. Soc. Am. Bull. 82, 563–572. petrophysics and flow modeling (Green Canyon 65, Gulf of Mexico). Mar. Petrol.
Dahlen, F.A., 1990. Critical taper model of fold-and-thrust belts and accretionary Geol. 21, 1311–1322.
wedges. Annu. Rev. Earth Planet. Sci. 18, 55–99. Fontanelli, P.D.R., De Ros, L.F., Remus, M.V.D., 2009. Provenance of deep-water reservoir
Davis, D.M., Engelder, T., 1985. The role of salt in fold-and-thrust belts. Tectonophysics sandstones from the Jubarte oil field, Campos Basin, Eastern Brazilian Margin. Mar.
119, 67–88. Petrol. Geol. 26, 1274–1298.
De Vera, J., Granado, P., McClay, K., 2010. Structural evolution of the Orange basin Forrest, J., Marcucci, E., Scott, P., 2005. Geothermal gradients and subsurface
gravity-driven system, offshore Namibia. Mar. Petrol. Geol. 27, 233–237. temperatures in the northern Gulf of Mexico. GCACS Trans. 55, 233–248.
DeJong, K.A., Scholten, R., 1973. Gravity and Tectonics. John Wiley, New York. 502 pp. Fowler, S.R., Mildenhall, J., Zalova, S., Riley, G., Elsley, G., Desplanques, A., Guliyev, F.,
DeMets, C., Gordon, R.G., Argus, D.F., Stein, S., 1990. Current plate motions. Geophys. J. 2000. Mud volcanoes and structural development on Shah Deniz. J. Pet. Sci. Eng. 28,
Int. 101, 425–478. 189–206.
Deptuck, M.E., Kendell, K., Smith, P., 2009. Complex deepwater fold-belts in the SW Franke, D., Barckhausen, U., Heyde, I., Tingay, M., Ramli, N., 2008. Seismic images of a
Sable Subbasin, offshore Nova Scotia. Frontiers Innovation, CSPG CSEG CWLS collision zone offshore NW Sabah/Borneo. Mar. Petrol. Geol. 25, 606–624.
Convention, Calgary, Alberta Canada. 4 pp. Frost, B.R., 1996. Structure and facies development in the Niger Delta resulting from
Deville, E., Battani, A., Griboulard, R., Guerlais, S., Herbin, J.P., Houzay, J.P., Muller, C., hydrocarbon maturation (abs.). AAPG Bull. 80, 1291.
Prinzhofer, A., 2003. The origin and processes of mud volcanism: new insights from Fruehn, J., White, R.S., Minshull, T.A., 1997. Internal deformation and compaction of the
Trinidad. In: Van Rensbergen, P., Hillis, R., Maltman, A.J., Morley, C.K. (Eds.), Makran accretionary wedge. Terra Nova 9, 101–104.
Subsurface Sediment Mobilization: Geol. Soc. Lond., Spec. Publ., 216, pp. 475–490. Galloway, W.E., Ganey-Curry, P.E., Xiang, L., Buffler, R.T., 2000. Cenozoic depositional
Deville, E., Guerlais, S.-H., Callec, Y., Griboulard, R., Huyghe, P., Lallemant, S., Mascle, A., history of the Gulf of Mexico Basin. AAPG Bull. 84, 1743–1774.
Noble, M., Schmitz, J., the collaboration of the Caramba working group, 2006. Gani, M.R., Alam, M.M., 2003. Sedimentation and basin-fill history of the Neogene
Liquefied vs stratified sediment mobilization processes: insight from the South of clastic succession exposed in the southeastern fold belt of the Bengal Basin,
the Barbados accretionary prism. Tectonophysics 428, 33–47. Bangladesh: a high-resolution sequence stratigraphic approach. Sed. Geol. 155,
Devlin, W.J., Cogswell, J.M., Gaskins, G.M., Isaksen, G.H., Pitcher, D.M., Puls, D.P., Stanley, 227–270.
K.O., Wall, G.R.T., 1999. South Caspian Basin – young, cool, and full of promise. GSA Garciacaro, E., Mann, P., Escalona, A., 2011. Regional structure and tectonic history of
Today 9, 1–9. the obliquely colliding Columbus foreland basin, offshore Trinidad and Venezuela.
Dickinson, W.R., 1982. Compositions of sandstones in circum-Pacific subduction Mar. Petrol. Geol. 28, 126–148.
complexes and fore-arc basins. AAPG Bull. 66, 121–137. Gardosh, M.A., Druckman, Y., 2006. Seismic stratigraphy, structure and tectonic
Dickinson, W.R., Suczek, C.A., 1979. Plate tectonics and sandstone composition. AAPG evolution of the Levant Basin, offshore Israel. In: Robertson, A.H.F., Mountrakis, D.
Bull. 63, 2164–2182. (Eds.), Tectonic Development of the Eastern Mediterranean Region: Geol. Soc.
Diegel, F.A., Karlo, J.F., Schuster, D.C., Shoup, R.C., Tauvers, P.R., 1995. Cenozoic structural Lond. Spec. Pub., 260, pp. 201–227.
evolution and tectono-stratigraphic framework of the northern Gulf Coast Garzanti, E., Critelli, S., Ingersoll, S.V., 1996. Paleogeographic and paleotectonic
continental margin. In: Jackson, M.P.A., Roberts, D.G., Snelson, S. (Eds.), Salt evolution of the Himalayan Range as reflected by detrital modes of Tertiary
tectonics a global perspective: AAPG Mem., 65, pp. 109–151. sandstones and modern sands (Indus transect, India and Pakistan). Geol. Soc. Am.
Dolan, P., Burggraf, D., Soofi, K., Fitzsimmons, R., Aydemir, E., Al, S.O., Strickland, L., 2004. Bull. 108, 631–642.
Challenges to exploration in frontier basins – the Barbados accretionary prism. Gay, A., Lopez, M., Cochonat, P., Séranne, M., Levaché, Sermondadaz, G., 2006. Isolated
AAPG International Conference, October 24–27, 2004, Cancun, Mexico. 6 pp. seafloor pockmarks linked to BSRs, fluid chimneys, polygonal faults and stacked
Dolson, J.C., Shann, M.V., Matbouly, S.I., Hammouda, H., Rashed, R.M., 2000. Egypt in the Oligocene-Miocene turbiditic palaeochannels in the Lower Congo Basin. Mar. Geol.
twenty-first century: petroleum potential in offshore trends. GeoArabia 6, 226, 25–40.
211–230. Gee, M.J.R., Uy, H.S., Warren, J., Morley, C.K., Lambiase, J.J., 2007. The Brunei slide: a giant
Doust, H., Omatsola, E., 1989. Niger Delta. In: Edwards, J.D., Santogrossi, P.A. (Eds.), submarine landslide on the North West Borneo Margin revealed by 3D seismic
Divergent/passive margin basins: AAPG Mem., 48, pp. 201–238. data. Mar. Geol. 246, 9–23.
88 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

Ghosh, A., Holt, W.E., Flesch, L.M., Haines, A.J., 2006. Gravitational potential energy of Housen, B.A., Tobin, H.J., Labaume, P., Leitch, E.C., Maltman, A.J., Ocean Drilling Program
the Tibetan Plateau and the forces driving the Indian Plate. Geology 34, 321–324. Leg 156 Shipboard Science Party, 1996. Strange decoupling across the decollement
Godon, A., Jandrzejewski, N., Castrec-Rouelle, M., Dia, A., Pineau, F., Boulègue, J., Javoy, of the Barbados accretionary prism. Geology 24, 127–130.
M., 2004. Origin and evolution of fluids from mud volcanoes in the Barbados Hudec, M.R., Jackson, M.P.A., 2004. Regional restoration across the Kwanza Basin,
accretionary complex. Geochim. Cosmochim. Acta 68, 2153–2165. Anglola: salt tectonics triggered by repeated uplift of a metastable passive margin.
Golonka, J., 2004. Plate tectonic evolution of the southern margin of Eurasia in the AAPG Bull. 88, 971–990.
Mesozoic and Cenozoic. Tectonophysics 381, 235–273. Hudec, M.R., Jackson, M.P.A., 2009. Interactions between spreading salt canopies and
Grando, G., McClay, K., 2004. Structural evolution of the Frampton growth fold system, their peripheral thrust systems. J. Struct. Geol. 31, 1114–1129.
Atwater Valley–Southern Green Canyon area, deep water Gulf of Mexico. Mar. Hung, J.-H., Wu, Y.-H., Yeh, E.-C., Wu, J.-C., Scientific Party, T.C.D.P., 2007. Subsurface
Petrol. Geol. 21, 889–910. structure, physical properties and fault zone characteristics in the scientific drill holes
Grando, G., McClay, K., 2007. Morphotectonics domains and structural styles in the of Taiwan Chelungpu-Fault Drilling Project. Terr. Atmos. Ocean. Sci. 18, 271–293.
Makran accretionary prism, offshore Iran. Sed. Geol. 196, 157–179. Hutchison, C.H., 1996. The ‘Rajang accretionary prism’ and ‘Lupar Line’ problem of
Graue, K., 2000. Mud volcanoes in deepwater Nigeria. Marine Petrol. Geol. vol. 17, Borneo. In: Hall, R., Blundell, D. (Eds.), Tectonic Evolution of Southeast Asia: Geol.
pp. 959–974. Soc. of Lond. Spec. Pub., 106, pp. 247–261.
Grigsby, J.D., Kassi, A.M., Khan, A.S., 2008. Diagenesis of the Oligocene-Early Miocene Hutchison, C.H., Bergman, S.C., Swauger, D.A., Graves, J.E., 2000. A Miocene collisional
Panjgur Formation, Paleocene Ispikan Formation and Exotic Blocks in the Makran belt in north Borneo: uplift mechanism and isostatic adjustment qualified by
Accretionary Belt, Southwest Pakistan: AAPG Annual Convention, San Antonio, thermochronology. J. Geol. Soc. Lond. 157, 783–794.
Texas. Huyghe, P., Mugnier, J.L., Griboulard, R., Deniaud, Y., Gonthier, E., Faugeres, J.C., 1999.
Guardado, A.R., Spadini, J.S., Brandao, L., Mello, R., 1989. Petroleum System of the Review of the tectonic controls and sedimentary patterns in late Neogene
Campos Basin, Brazil. In: Edwards, J.D., Santogrossi, P.A. (Eds.), Divergent/passive piggyback basins on the Barbados Ridge Complex. In: Mann, P. (Ed.), Caribbean
margin basins: AAPG Mem., 48, pp. 317–324. basins. Elsevier, Amsterdam, pp. 369–388.
Guiraud, R., Issawi, B., Bosworth, W., 2001. Phanerozoic history of Egypt and Huyghe, P., Foata, M., Deville, E., Mascle, G., the Caramba Working Group, 2004. Channel
surrounding areas. In: Ziegler, P.A., Cavazza, W., Robertson, A.H.F., Crasquin-Soleau, profiles through the active thrust front of the southern Barbados prism. Geology 32,
S. (Eds.), Peri-Tethys Mem. 6: Peri-Tethyan Rift/Wrench Basins and Passive 429–432.
Margins: Mém. Mus. natn. Hist. nat., 186, pp. 469–509. Ingram, G.M., Chisholm, T.J., Grant, C.J., Hedlund, C.A., Stuart-Smith, P., Teasdale, J.,
Guliev, I.S., 1992. A review of mud volcanism. Translation of the report by: Azerbaijan 2004. Deepwater North West Borneo: hydrocarbon accumulation in an active fold
Academy of Sciences Institute of Geology 65 pp. and thrust belt. Mar. Petrol. Geol. 21, 879–887.
Haack, R.C., Sundararaman, P., Diedjomahor, J.O., Xiao, H., Grant, N.J., May, E.D., Kelsch, Ings, S.J., Beaumont, C., 2010. Continental margin shale tectonics: preliminary results
K., 2000. Niger Delta petroleum systems, Nigeria. In: Mello, M.R., Katz, B.J. (Eds.), from coupled fluid-mechanical models of basin scale delta instability. J. Geol. Soc.
Petroleum Systems of South Atlantic margins: AAPG Mem., 73, pp. 213–232. 167, 571–582.
Hafkenscheid, E., Wortel, M.J.R., Spakman, W., 2006. Subduction history of the Tethyan Ings, S.J., Shimeld, J.W., 2006. A new conceptual model for the structural evolution of a
region derived from seismic tomography and tectonic reconstructions. J. Geophys. regional salt detachment on the northeast Scotian margin, offshore eastern Canada.
Res. 111. doi:10.1029/2005JB003791. AAPG Bull. 90, 1407–1423.
Hall, R., 2002. Cenozoic geological and Plate Tectonic evolution of SE Asia and the SW Jackson, M.P.A., Vendeville, B.C., Schultz-Ela, D.D., 1994. Structural dynamics of salt
Pacific: computer-based reconstructions, model and animations. J. Asian Earth Sci. systems. Annu. Rev. Earth Planet. Sci. 22, 93–117.
20, 353–434. Jackson, J., Priestley, K., Allen, M., Berberian, M., 2002. Active tectonics of the South
Hall, R., 2009. The Eurasian SE Asian margin as a modern example of an accretionary Caspian Basin. Geophys. J. Int. 148, 214–245.
orogen. In: Cawood, P.A., Kroner, A. (Eds.), Earth Accretionary Systems in Space and Jackson, M.P.A., Hudec, M.R., Jennette, D.C., Kilby, R.E., 2008. Evolution of the Cretaceous Astrid
Time: Geol. Soc. Lond. Spec. Pub., 318, pp. 351–372. thrust belt in the ultradeep-water Lower Congo Basin, Gabon. AAPG Bull. 92, 487–511.
Hall, R., in press, Australia-SE Asia collision: plate tectonics and crustal flow. In: R. Hall, James, D.M.D., 1984. The Geology and Hydrocarbon Resources of Negara Brunei
M.A. Cottam, M.J.E. Wilson, (editors), The SE Asian gateway: history and tectonics Darussalam. Spec. Pub., Muzium Brunei and Brunei Shell Petroleum Company
of Australia-Asia collision. Geol. Soc. Lond. Spec. Pub. Berhad. 164 pp.
Hall, R., Morley, C.K., 2004. Sundaland Basins. AGU Geophys. monogr. 149, 55–85. Jolivet, L., Augier, R., Robin, C., Suc, J.P., Rouchy, J.M., 2006. Lithosphere-scale
Hall, R., Nichols, G.J., 2002. Cenozoic sedimentation and tectonics in Borneo: climatic geodynamic context of the Messinian salinity crisis. Sed. Geol. 188–189, 9–33.
influences on orogenesis. In: Jones, S.J., Frostick, L. (Eds.), Sediment Flux to Basins, Kadarusman, A., Miyashita, S., Maruymama, S., Parkinson, C.D., Ishikawa, A., 2004.
Causes, Controls and Consequences: Geol. Soc. Lond. Spec. Pub., 191, pp. 5–22. Petrology, geochemistry and paleogeographic reconstruction of the East Sualwesi
Hall, R., Wilson, M.E.J., 2000. Neogene sutures in eastern Indonesia. J. Asian Earth Sci. 18, Ophiolite, Indonesia. Tectonophysics 392, 55–83.
781–808. Kaluza, M.J., Doyle, E.H., 1996. Detecting fluid migration in shallow sediments: continental
Hall, R., Cloke, I., Nur'aini, S., Puspita, S.D., Calvert, S.J., Elders, C., 2009. The North slope environment, Gulf of Mexico. In: Schumacher, D., Abrams, M.A. (Eds.),
Makassar Straits: what lies beneath? Pet. Geosci. 15, 147–158. Hydrocarbon migration and its near-surface expression: AAPG Mem., 66, pp. 15–26.
Hamilton, R., De Vera, J., 2009. A review and global comparison of deepwater fold and Karig, D.E., Barber, A.J., Charlton, T.R., Klemperer, S., Hussong, D.M., 1987. Nature and
thrust belt settings – implications for their hydrocarbon prospectivity. Shell distribution of deformation across the Banda Arc–Australian collision zone at
University Lecture Series, Geological Society of London. http://www2.geolsoc.org. Timor. Geol. Soc. Am. Bull. 98, 18–32.
uk/presentations/090311Hamilton/#. Katz, B.J., 2003. Hydrocarbon charge in deepwater Nigeria – an evolving story. 2003
Hardman, R.F.P., 1982. Chalk reservoirs of the North Sea. Bull. Geol. Soc. Denmark 30, AAPG International Conference & Exhibition Technical Program, Abstract.
119–137. Katz, B.J., Mello, M., 2000. Petroleum systems of South Atlantic margin basins – an
Harms, J.C., Cappel, H.N., Francis, D.C., 1985. The Makran Coast of Pakistan: its overview. In: Mello, M.R., Katz, B.J. (Eds.), Petroleum Systems of South Atlantic
stratigraphy and hydrocarbon potential. In: Haq, B.U. and Milliman, J.D. (Eds.), margins: AAPG Mem., 73, pp. 1–14.
Marine geology and oceanography of Arabian Sea and coastal Pakistan, New York, Katz, B., Richards, D., Long, D., Lawrence, W., 2000. A new look at the components of the
Van Nostrand Reinhold, pp. 3–26. petroleum system of the South Caspian Basin. J. Pet. Sci. Eng. 28, 161–182.
Helset, H.M., Lander, R.H., Matthews, J.C., Reemst, P., Bonnell, L.M., Frette, I., 2002. The Kaul, A., Rosenberger, A., Villinger, H., 2000. Comparisons of measured and BSR-derived
role of diagenesis in the formation of fluid overpressures in clastic rocks. Nor. heat flow values, Makran accretionary prism, Pakistan. Mar. Geol. 164, 37–51.
Petrol. Soc. Spec. Pub. 11, 37–50. Kehle, R.O., 1989. The origin of salt structures. In: Schreiber, C.B. (Ed.), Evaporites and
Henry, P., Jouniaux, L., Screaton, E.J., Hunze, S., Saffer, D.M., 2003. Anisotropy of hydrocarbons. Columbia Univ. Press, NY, pp. 345–404.
electrical conductivity record of initial strain at the toe of the Nankai accretionary Khan, M.A., Raza, H.A., Alam, S., 2007. Petroleum geology of the Makran region:
prism. J. Geophys. Res. 108. doi:10.1029/2002JB002287. implications for hydrocarbon occurrence in cool basins. J. Pet. Geol. 14, 5–18.
Hesse, S., Back, S., Franke, D., 2008. The deepwater fold-thrust belt offshore NW Borneo: King, R.C., Hillis, R.R., Tingay, M.R.P., Morley, C.K., 2009. Present-day stress and
gravity-driven versus basement-driven shortening. Geol. Soc. Am. Bull. 121, neotectonic provinces of the Baram Delta and deep-water fold-thrust belt. J. Geol.
939–953. Soc. Lond. 166, 1–4.
Hesse, S., Back, S., Franke, D., 2010. The structural evolution of folds in a deepwater fold King, R., Backé, G., Morley, C.K., Hillis, R., Tingay, M., 2010. Balancing deformation in NW
and thrust belt a case study from the Sabah continental margin, offshore NW Borneo: quantifying plate-scale vs gravitational tectonics in a delta and deepwater
Borneo, SE Asia. Mar. Petrol. Geol. 27, 442–454. fold-thrust belt system. Mar. Petrol. Geol. 27, 238–246.
Hill, R.J., Schenk, C.J., 2005. Petroleum geochemistry of oil and gas from Barbados: King, R., Tingay, M.R.P., Hillis, R.R., Morley, C.K., Clark, J., in press. Present-day stress
implications for distribution of Cretaceous source rocks and regional petroleum orientation across the NW Borneo collisional margin. J. Geophys. Res.
prospectivity. Mar. Petrol. Geol. 22, 917–943. Knapp, C.C., Knapp, J.H., Connor, J.A., 2004. Crustal-scale structure of the South Caspian
Hinz, K., Fritsch, J., Kempter, E.H.K., Mohammad, A.M., Meyer, J., Mohamed, D., Vosberg, Basin revealed by deep seismic reflection profiling. Mar. Petrol. Geol. 21, 1073–1081.
H., Weber, J., Benavidez, J., 1989. Thrust tectonics along the north-western Knox, G.J., Omatsola, E.M., 1989. Development of the Cenozoic Niger Delta in terms of
continental margin of Sabah/Borneo. Geol. Rundsch. 78, 705–730. the ‘Escalator Regression’ model and impact on hydrocarbon distribution. In: van
Hossain, M.S., Uddin, A., Savrda, M.-K., Hames, C.E., Uddin, Z., Iman, M.B., 2009. der Linden, W.J.M., Cloetingh, S.A.P.L., Kaasschieter, J.P.K., van der Gun, J.A.M. (Eds.),
Distribution of overpressure in the lower Miocene Bhuban Formation in the Bengal Proceedings KNGMG symposiums coastal lowlands, geology and geotechnology.
basin; Abstracts with Programs. Geol. Soc. Am. 41, 56. Kluwer Academic Publishers, Amsterdam, pp. 181–202.
Hosseini-Barzi, M., Talbot, C.J., 2003. A tectonic pulse in the Makran accretionary prism Kopf, A., Brown, K.M., 2003. Friction experiments on saturated sediments and their
recorded in Iranian coastal sediments. J. Geol. Soc. Lond. 160, 903–910. implications for the stress state of the Nankai and Barbados subduction thrusts.
Housen, B.A., 1997. Magnetic anisotropy of Barbados Prism Sediments. In: Shipley, Mar. Geol. 202, 193–210.
T.H., Ogawa, Y., Blum, P., Bahr, J.M. (Eds.), Proceed. Ocean. Drill. Prog. Sci. Res., Kopf, A., Deyhle, A., Lavrushin, V.Y., Polyak, B.G., Gieskes, J.K., Buachidze, G.I., Wallman,
156, pp. 97–105. K., Eisenhauer, A., 2003. Isotopic evidence (He, B, C) for deep fluid and mud
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 89

mobilization from mud volcanoes in the Caucasus continental collision zone. Int. J. Meyer, D., Zarra, L., Rains, D., Meltz, B., Hall, T., 2005. Emergence of the Lower Tertiary
Earth Sci. 92, 407–425. Wilcox Trend in the Deepwater Gulf of Mexico: World Oil, pp. 72–77. May.
Kostenko, O.V., Naruk, S.J., Hack, W., Poupon, M., Meyer, H.-J., Mora-Glukstad, M., Meyer, D., Zarra, L., Yun, J., 2007. From BAHA to Jack, Evolution of the Lower Tertiary
Anowai, C., Mordi, M., 2008. Structural evaluation of column-height controls at a Wilcox Trend in the Deepwater Gulf of Mexico. Sed. Record 5, 4–9.
toe-thrust discovery, deep-water Niger Delta. AAPG Bull. 92, 1615–1638. Michard, A., Negro, F., Frizon de Lamotte, D., Saddiqi, O., 2007. Serpentinite slivers and
Krueger, A., Gilbert, E., 2009. Deepwater fold-thrust belts: not all the beasts are equal. metamorphism in the external Maghrebides: arguments for an intracontinental suture
AAPG Search and Discovery Article #30085. http://www.searchanddiscovery.com/ in the African paleomargin (Morocco, Algeria). Rev. Soc. Geol. Espana 20, 173–185.
documents/2009/30085krueger/index.htm. Milton, C.J., 2006. The Kikeh Field, Sabah, East Malaysia. AAPG International Conference
Lawrence, S., Babaev, H., 2000. Large structures indicated off Turkmenistan. Oil Gas J. 98, and Exhibition, November 5–8, Technical Program, Perth, Australia.
86–89. Mitchell, A.G.H., 1993. Cretaceous–Cenozoic tectonic events in the Western Burma
Letouzey, J., Colletta, B., Vially, R., Chermette, J.C., 1995. Evolution of salt related (Burma)–Assam region. J. Geol. Soc. (Lond.) 150, 1089–1102.
structures in compressional setting. In: Jackson, M.P.A., Roberts, D.G., Snelson, S. Moore, J.C., Tobin, H., 1997. Estimate fluid pressures of the Barbados accretionary prism
(Eds.), Salt tectonics: a global perspective: AAPG mem., 65, pp. 41–60. and adjacent sediments. Proc. Ocean Drill. Prog. Sci. Res. 156, 229–239.
Levell, B.K., 1987. The nature and significance of regional unconformities in the Moore, J.C., Mascle, A., et al., 1990. Proceedings of the Ocean Drilling Program, Scientific
hydrocarbon-bearing Neogene sequences offshore West Sabah. Geol. Soc. Malay. results. College Station, Texas, Ocean Drilling Program, Volume 110. 448 pp.
Bull. 21, 55–90. Moore, J.C., et al., 1995. Abnormal fluid pressures and fault-zone dilation in the Barbados
Lewis, S.D., Ladd, J.W., Bruns, T.R., 1988. Structural development of an accretionary accretionary prism: evidence from logging while drilling. Geology 23, 605–608.
prism by thrust and strike-slip faulting: Shumagin region, Aleutian Trench. Geol. Moore, J.C., Rowe, C., Meneghini, F., 2007. How accretionary prisms elucidate
Soc. Am. Bull. 100, 767–782. seismogenesis in subduction zones. In: Dixon, T.H., Moore, J.C. (Eds.), The
Linthout, K., Helmers, H., Wijbrans, J.R., Van Wees, J.D., 1996. 40Ar/39Ar constraints on seismogenic zone of subduction thrust faults. Columbia University Press, New
obduction of the Seram ultramafic complex: consequences for the evolution of the York, pp. 288–313.
southern Banda Sea. In: Hall, R., Blundell, D. (Eds.), Tectonic Evolution of Southeast Morgan, R., 2003. Prospectivity in ultradeep water: the case for petroleum generation
Asia: Geol. Soc. Spec. Pub., 106, pp. 455–464. and migration within the outer parts of the Niger Delta apron. In: Arthur, T.J.,
Loncke, L., Gaullier, V., Mascle, J., Vendeville, B., Camera, L., 2006. The Nile deep-sea fan: McGregor, D.S., Cameron, N.R. (Eds.), Petroleum Geology of Africa: new themes and
an example of interacting sedimentation, salt tectonics, and inherited subsalt developing technologies: Geol. Soc. Lond. Spec. Pub., 207, pp. 151–164.
paleotopographic features. Mar. Petrol. Geol. 23, 297–315. Morgan, J.K., Karig, D.E., 1995. Kinematics and a balanced and restored cross-section
Louden, K.E., Mareschal, J.-C., 1996. Measurements of radiogenic heat production on across the toe of the eastern Nankai accretionary prism. J. Struct. Geol. 17, 31–45.
basement samples from sites 897 and 900. In: Whitmarsh, R.B., Sawyer, D.S., Klaus, Morgan, J.P., Coleman, J.M., Gagliano, S.M., Morgan, J.P., Coleman, J.M., Gagliano, S.M.,
A., Masson, D.G. (Eds.), Proc. Ocean Drill. Prog., Sci. Res., 149, pp. 675–682. 1968. Mud-lumps: diapiric structures. In: Braunstein, J., O'Brien, G.D. (Eds.), AAPG.
MacKay, M.E., 1995. Structural variation and landward vergence at the toe of the Mem., 8, pp. 145–161.
Oregon accretionary prism. Tectonics 14, 1309–1320. Morley, C.K., 1988. Out-of-sequence thrusts. Tectonics 7, 539–561.
MacKay, M.E., Moore, G.F., Cochrane, G.R., Moore, J.C., Kulm, L.D., 1992. Landward Morley, C.K., 1994. Fold-generated imbricates: examples from the Caledonides of
vergence and oblique subduction trends in the Oregon margin accretionary prism: Southern Norway. J. Struct. Geol. 16, 619–631.
implications and effect on fluid flow. Earth Planet. Sci. Lett. 109, 477–491. Morley, C.K., 2003a. Mobile shale related deformation in large deltas developed on passive
Madon, M., Meng, L.K., Anuar, A., 1999. Chapter 22, Sabah Basin. The Petroleum Geology and active margins: review and new views. In: Van Rensbergen, P.R., Hillis, R.,
and Resources of Malaysia, Petronas, Kuala Lumpur, Malaysia, pp. 501–542. Maltman, A., Morley, C.K. (Eds.), Subsurface Sediment Mobilization: Geol. Soc. Lond.
Maksoud, J., 2008. Deepwater field onshore in Nigeria goes onstream: E&P Mag. http:// Spec. Pub., 216, pp. 335–357.
www.epmag.com/webonly/item7294.php. Morley, C.K., 2003b. Outcrop examples of mudstone intrusions from the Jerudong
Maltman, A.J., Byrne, T., Karig, D.E., Lallemant, S.J., Knipe, R., Prior, D., 1993. Deformation anticline, Brunei Darussalam, and inferences for hydrocarbon reservoirs. In: Van
structures at Site 808, Nankai accretionary prism, Japan. Rensbergen, P.R., Hillis, R., Maltman, A., Morley, C.K. (Eds.), Subsurface Sediment
Maltman, A., Labaume, P., Housen, B., 1997. Structural geology of the decollement at the Mobilization: Geol. Soc. Lond. Spec. Pub., 216, pp. 381–394.
toe of the Barbados Accretionary Prism. In: Shipley, T.H., Ogawa, Y., Blum, P., Bahr, Morley, C.K., 2007a. Interaction between critical wedge geometry and sediment supply
M.J. (Eds.), Proc. Ocean Drill. Prog., Sci. Results, vol. 156, pp. 279–292. in a deepwater fold belt, NW Borneo. Geology 35, 139–142.
Mandl, G.W., 1999. Short notes on the Hallstatt salt rock – the “Haselgebirge”. Ber. d. Morley, C.K., 2007b. Development of crestal normal faults associated with deepwater
Geol. Bund. 49, 91–95. fold growth. J. Struct. Geol. 29, 1148–1163.
Mann, D., Kukowski, N., 1998. Numerical modeling of focused fluid flow in the Cascadia Morley, C.K., 2009a. Growth of folds in a deep-water setting. Geosphere 5, 59–89.
accretionary wedge. J. Geodyn. 27, 359–372. Morley, C.K., 2009b. Geometry of an oblique thrust fault zone in a deepwater fold belt
Marsaglia, K.M., Barragan, J.G.C., Padilla, I., Milliken, K.L., 1996. 11 evolution of the from 3D seismic data. J. Struct. Geol. 31, 1540–1555.
Iberian Passive Margin as reflected in sand provenance. In: Whitmarsh, R.B., Morley, C.K., Back, S., 2008. Estimating hinterland exhumation from late orogenic basin
Sawyer, D.S., Klaus, A., Masson, D.G. (Eds.), Proc. Ocean Drilling Prog. Sci. Res., 149, volume, NW Borneo. J. Geol. Soc. Lond. 165, 353–366.
pp. 269–280. Morley, C.K., Guerin, G., 1996. Comparison of gravity-driven deformation styles and
Marton, L.G., Tari, G.C., Lehmann, C.T., 2000. Evolution of the Angolan passive margin, West behaviour associated with mobile shales and salt. Tectonics 15, 1154–1170.
Africa, with emphasis on post-salt structural styles. In: Mohriak, W., Talwani, M. (Eds.), Morley, C.K., Leong, L.C., 2008. Evolution of deep-water synkinematic sedimentation in a
Atlantic rifts and continental margins: Am. Geophys. Union Geophys. Mono., 115, piggyback basin, determined from three-dimensional seismic data. Geosphere 4,
pp. 1219–1230. 939–962.
Mascle, A., Endignoux, L., Chennouf, T., 1990. Frontal accretion and piggyback basin Morley, C.K., Westaway, R., 2006. Subsidence in the super-deep Pattani and Malay
development at the southern edge of the Barbados Ridge accretionary complex. In: basins of Southeast Asia: a coupled model incorporating lower-crustal flow in
Moore, J.C., Mascle, A. (Eds.), Proc. Ocean Drill. Prog., Sci. Res., 110, pp. 17–28. response to post-rift sediment loading. Basin Res. 18, 51–84.
Masson, D.G., Milsom, J., Barber, A.J., Sikumbang, N., Dwiyanto, B., 1991. Recent Morley, C.K., Crevello, P., Ahmad, Zulkifli, 1998. Shale tectonics-deformation associated with
tectonics around the island of Timor, eastern Indonesia:. Mar. Petrol. Geol. 8, 35–49. active diapirism: the Jerudong Anticline, Brunei Darussalam. J. Geol. Soc. Lond. 155,
Maudit, T., Guerin, G., Brun, J.-P., Lecanu, H., 1997. Raft tectonics: the effects of basal 475–490.
slope angle and sedimentation rate on progressive extension. J. Struct. Geol. 19, Morley, C.K., Back, S., Crevello, P., Van Rensbergen, P., Lambiase, J.J., 2003. Characteristics of
1219–1230. repeated, detached, Miocene–Pliocene tectonic inversion events, in a large delta province
Maurin, T., Rangin, C., 2009. Structure and kinematics of the Indo-Burmese Wedge: on an active margin, Brunei Darussalam, Borneo. J. Struct. Geol. 25, 1147–1169.
recent and fast growth of the outer wedge. Tectonics 28. doi:10.1029/ Morley, C.K., Tingay, M., Hillis, R., King, R., 2008. Relationship between structural
2008TC002276. style, overpressures, and modern stress, Baram Delta Province, northwest Borneo.
McClay, K.R., Dooley, T., Lewis, G., 1998. Analog modeling of progradational delta J. Geophys. Res. 113. doi:10.1029/2007JB005324.
systems. Geology 26, 771–774. Mostofi, B., Gansser, A., 1957. The story behind the 5 Alborz. Oil Gas J. 55, 78–84.
McClay, K., Munoz, J.-A., Garcia-Senz, J., 2004. Extensional salt tectonics in a Nagihara, S., Opre Jones, K., 2005. Geothermal heat flow in the northeast margin of the
contractional orogen a newly identified event in the Spanish Pyrenees. Geology Gulf of Mexico. AAPG Bull. 89, 821–831.
32, 737–740. Nielsen, C., Chamote-Rooke, N., Rangin, C., the ANDAMAN Cruise Team, 2004. From
McConnell, D.R., Kendall, B.A., 2003. Images of the base of gas hydrate stability in the partial to full strain partitioning along the Indo-Burmese hyper-oblique subduc-
deepwater Gulf of Mexico. Lead. Edge 22, 361–367. tion. Mar. Geol. 209, 303–327.
McConnell, D., Gharib, J., Henderson, J., Danque, H.-W.A., Digby, A., 2008. Multibeam Nikishin, A.M., Ziegler, P.A., Panov, D.I., Nazarevich, B.P., Brunet, M.-F., Stephenson, R.A.,
echosounders and seismic data detect seeps, enhancing hydrocarbon exploration Boloyov, S.N., Korotaev, M.V., Tikhomirov, P.L., 2001. Mesozoic and Cainozoic
potential: World Oil Magazine, 229. 10 pp. http://www.tdi-bi.com/our_publica- evolution of the Scythian Platform-Black Sea-Caucasus domain. In: Ziegler, P.A.,
tions/our_pubs_main.htm. Cavazza, W., Robertson, A.H.F., Crasquin-Soleau, S. (Eds.), Peri-Tethys Mem. 6: Peri-
McGilvery, T.A., Cook, D.L., 2003. The influence of local gradients on accommodation Tehtyan Rift-Wrench Basins and Passive Margins: Mem. Mus. Natn. Hist. nat. Paris,
space and linked depositional elements across a stepped slope profile, offshore 186, pp. 295–346.
Brunei. In: Roberts, H., Rose, N., Fillon, R.H., Anderson, J.B. (Eds.), Shelf margin Nilandaroe, N., Mogg, W., Barraclough, R., 2002. Characteristics of the fractured
deltas and linked downslope petroleum systems: Global significance and future carbonate reservoir of the Oseil field, Seram island, Indonesia. Proc. Annu. Conv.
exploration potential: Houston, Texas, Gulf Coast Section Society of Economic Indonesian Petrol. Assoc. 28 (1), 439–456.
Paleontologists and Sedimentologists, 23rd Annual Bob F. Perkins Research Nixon, L.D., Shepard, N.K., Bohannon, C.M., Montgomery, T.M., Kazanis, E.G., Gravois, M.P.,
Conference, pp. 387–419. 2009. Deepwater Gulf of Mexico 2009: Interim Report of 2008 Highlights: U.S.
Metevier, F., Gaudemer, Y., Tapponnier, P., Klein, M., 1999. Mass accumulation rates in Department of the Interior Minerals Management Service, Gulf of Mexico OCS Region,
Asia during the Cenozoic. Geophys. J. Int. 137, 280–318. New Orleans, MMS 2009-016. 72 pp.
90 C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91

O'Brien, G.W., Lisk, M., Duddy, I.R., Hamilton, P.J., Woods, P., Cowley, R., 1999. Plate Searle, M.P., Stevens, R.K., 1984. Ophiolite emplacement and obduction. Geol. Soc. Lond.
convergence, foreland development and fault reactivation: primary controls on Spec. Pub. 13, 303–319.
brine migration, thermal histories and trap breach in the Timor Sea Australia. Mar. Shipley, T.H., Moore, G.F., Bangs, N.L., Moore, J.C., Stoffa, P.L., 1994. Seismically inferred
Petrol. Geol. 16, 533–560. dilatancy distribution, northern Barbados Ridge decollement: Implications for fluid
Ofurhie, M.A., Agha, G.U., Lufadeju, A.O., Ineh, G.C., 2002. Turbidite depositional migration and fault strength. Geology 22, 411–414.
environment in deepwater of Nigeria. 2002 Offshore Technology Conference, Shipley, T.H., Ogawa, J.J., Blum, P. (Eds.), 1995. Proceedings of the Ocean Drilling
Houston, Texas, 6–9 May, 2002, OTC 14068. 9 pp. Program, Initial Results: Texas College Station, 156, pp. 13–27.
Osborne, M.J., Swarbrick, R.E., 1998. Mechanisms for generating overpressure in Sikder, A.M., Alam, M.M., 2003. 2-D modeling of the anticlinal structures and structural
sedimentary basins: a reevaluation. AAPG Bull. 81, 1023–1041. development of the eastern fold belt of the Bengal Basin, Bangladesh. Sed. Geol.
Osborne, M.J., Swarbrick, R.E., 1999. Diagenesis in North Sea HPHT clastic reservoirs- 155, 209–226.
consequences for porosity and overpressure prediction. Mar. Petrol. Geol. 16, Simpson, G.D.H., 2009. Mechanical modelling of folding versus faulting in brittle-ductile
337–353. wedges. J. Struct. Geol. 31, 369–381.
Pairault, A.A., Hall, R., Elders, C.F., 2003. Structural styles and tectonic evolution of the Sone, H., Yeh, E.-C., Nakaya, T., Hung, J.-H., Ma, K.-F., Wang, C.-Y., Song, S.-R.,
Seram Trough, Indonesia. Mar. Petrol. Geol. 20, 1141–1160. Shimamoto, T., 2007. Mesoscopic structural observations of cores from the
Parkinson, C., 1998. Emplacement of the east Sulawesi ophiolite: evidence form Chelungup Fault System, Taiwan Chelungpu-Fault Drilling Project, hole-A, Taiwan.
subophiolitic rocks. J. Asian Earth Sci. 16, 13–28. Terr. Atmos. Ocean. Sci. 18, 359–377.
Peel, E.J., Hossack, J.R., Travis, C.J., 1995. Genetic structural provinces and salt tectonics of Spathopoulos, F., 1996. An insight on salt tectonics in the Angola Basin, South Atlantic.
the Cenozoic offshore U.S. Gulf of Mexico: a preliminary analysis. In: Jackson, M.P.A., In: Alsop, G.I., Blundell, D.J., Davison, I. (Eds.), salt tectonics: Geol. Soc. Lond. Spec.
Roberts, D.G., Snelson, S. (Eds.), Salt Tectonics: A Global Perspective: AAPG Mem., 65, Pub., 100, pp. 153–174.
pp. 153–175. Spötl, C., 1989. The Alpine Haselgebirge Formation, Northern Calcareous Alps (Austria):
Perez, O., Bilham, R., Bendick, R., Velandia, J., Hernadez, C., Hoyer, M., Kozuch, M., 2001. Permo-Scythian evaporites in an Alpine thrust system. Sed. Geol. 65, 113–125.
Velocity field across the southern Caribbean plate boundary and estimates of Stampfli, G.M., Borel, G.D., 2004. The TRANSMED Transects in Space and Time:
Caribbean-South American plate motion using GPS geodesy 1994–2000. Geo. Res. Constraints on the Paleotectonic Evolution of the Mediterranean Domain. In:
Lett. 28, 2987–2990. Cavazza, W., Roure, F.M., Spakma, W., Stampfli, G.M., Ziegler, P.A. (Eds.), The
Pindell, J.L., Kennan, L., 2007. Cenozoic kinematics and dynamics of oblique collision TRANSMED Atlas - The Mediterranean Region from Crust to Mantle. Springer,
between two convergent plate margins: the Caribbean-South America Collision in Berlin, pp. 53–80. Compact Disc CD-Rom.
Eastern Venezuela, Trinidad and Barbados. Trans. GCSSEPM, 27th Annual Bob F. Standley, C.E., Harris, R., 2009. Tectonic evolution of forearc nappes of the active Banda
Perkins Research Conference, pp. 458–553. arc-continent collision: Origin, age metamorphic history and structure of the
Platt, J.P., 1986. Dynamics of orogenic wedges and the uplift of high-pressure Lolotoi Complex, East Timor. Tectonophysics 479, 66–94.
metamorphic rocks. Geol. Soc. Am. Bull. 97, 1037–1053. Stearns, D.W., Friedman, M., 1972. Reservoirs in fractured rock. AAPG Mem. 16, 82–106.
Platt, J.P., Leggett, J.K., Young, J., Raza, H., Alam, S., 1985. Large-scale sediment underplating Steckler, M.S., Akhter, S.H., Seeber, L., 2008. Collision of the Ganges-Brahmaputra Delta
in the Makran accretionary prism, southwest Pakistan. Geology 13, 507–511. with the Burma Arc: Implications for earthquake hazard. Earth Planet. Sci. Lett. 273,
Platt, J.P., Leggett, J.K., Alam, S., 1988. Slip vectors and fault mechanics in the Makran 367–378.
accretionary wedge, southwest Pakistan. J. Geophys. Res. 93, 7955–7973. Stewart, S.A., Davies, R.J., 2006. Structure and emplacement of mud volcano systems in
Rajnauth, J., Vincent, H., Roberts, C., 2004. Trinidad and Tobago: The Atlantic Basin the South Caspian Basin. AAPG Bull. 90, 771–786.
(deep water) challenge. SPE Annual Technical Conference and Exhibition, 26-29th Stork, A.L., Selby, N.D., Heyburn, R., Searle, M.P., 2008. Accurate relative earthquake
September 2004, Houston, Texas. doi:10.2118/90809-MS. hypocenters reveal structure of the Burmese subduction zone. Bull. Seis. Soc. Am.
Ramberg, H., 1981. Gravity, deformation and the Earth's crust in theory, experiments 98, 2815–2827.
and geological application (2nd Edition). Academic Press, London. 452 pp. Suppe, J., 1983. Geometry and kinematics of fault-bend folding. Am. J. Sci. 283, 684–721.
Robertson, A.H.F., Mountrakis, D., 2006. Tectonic development of the Eastern Suppe, J., 2007. Absolute fault and crustal strength from wedge tapers. Geology 35,
Mediterranean region: an introduction. In: Robertson, A.H.F., Mountrakis, D. 1127–1130.
(Eds.), Tectonic development of the Eastern Mediterranean region: Geol. Soc. Lond. Suppe, J., Medwedeff, D.A., 1990. Geometry and kinematics of fault-propagation folding.
Spec. Pub., 260, pp. 1–9. Ecl. Geol. Helv. 83, 409–454.
Roure, F., Nazaj, S., Mushka, K., Fili, I., Cadet, J.-P., Bonneau, M., 2004. Kinematic Suppe, J., Connors, C.D., Zhang, Y., 2004. Shear fault-bend folding. In: McClay, K.R. (Ed.),
evolution and petroleum systems- an appraisal of the outer Albanides. In: McClay, Thrust tectonics and hydrocarbon systems: AAPG Mem., 82, pp. 303–323.
K.R. (Ed.), Thrust tectonics and hydrocarbon systems: AAPG Mem., 82, pp. 472–493. Suter, G., 1980a, Carte geologique du Rif, 1/500,000: Notes et Mem. Serv. Geol. Maroc,
Rowan, M.G., Peel, F.J., Vendeville, B.C., 2004. Gravity-driven fold belts on passive 245a.
margins. In: McClay, K.R. (Ed.), Thrust tectonics and hydrocarbon systems: AAPG Suter, G., 1980b. Carte structural du Rif, 1/500,000. Notes et Mem. Serv. Geol. Maroc,
Mem., 82, pp. 157–182. 245b.
Ruarri, D.-S., McDonnell, A., Wood, L., 2009. Characteristics of mobile shale in the deep Tan, D.N.K., Lamy, J.M., 1990. Tectonic evolution of the NW Sabah continental margin
stratigraphic subsurface. AAPG Search and Discovery Article 90090, 2009 AAPG since the Late Eocene. Bull. Geol. Soc. Malay. 27, 241–260.
Annual Convention and Exhibition, Denver Colorado, June 7-10. Tari, G., Molnar, J., Ashton, P., 2003. Examples of salt tectonics from West Africa: a
Saleh, J., Edwards, K., Barbaste, J., Balkaransingh, S., Grant, D., Weber, J., Leong, T., 2004. comparative approach. In: Doyle, P., Gregory, F.J., Griffiths, J.S., Hartley, A.J.,
On some improvements in the geodetic framework of Trinidad and Tobago. Surv. Holdsworth, R.E., Morton, A.C., Robins, N.S., Stocker, M.S., Turner, J.P. (Eds.),
Rev. 37, 604–625. Petroleum Geology of Africa: New Themes and Developing Technologies: Geol. Soc.
Saller, A.H., Lin, R., Dunham, J.B., 2006. Leaves in turbidite sands: The main source of oil Lond. Spec. Publ., 207, pp. 85–104.
and gas in the deep-water Kutei Basin, Indonesia. AAPG Bull. 90, 1585–1608. Taylor, E.P., Burkett, J., Wackler, J.D., Leonard, J.N., 1990. Chapter 22, Physical properties and
Saller, A., Werner, K., Sugiaman, F., Cebastiant, A., May, R., Glenn, D., Barker, C., 2008. microstructural response of sediments to accretion-subduction: Barbados.Forearc. In:
Characteristics of Pleistocene deep-water fan lobes and their application to an Bennett, R.H., Bryant, W.R., Hulbert, M.H. (Eds.), The microstructure of fine-grained
upper Miocene reservoir model, offshore East Kalimantan, Idonesia. AAPG Bull. 92, sediments, from mud to shale. Springer-Verlag, New York, pp. 213–228.
919–949. Teisserenc, P., Villemin, J., 1989. Sedimentary basins of Gabon – geology and oil systems.
Samsu, D.H., Darma, W.I., Meisling, K., Boyd, D.T., Hertig, S., Yough, C.T., 2000. Deep In: Edwards, J.D., Santogrossi, P.A. (Eds.), Divergent/passive margin basins: AAPG
Water Exploration Potential of the Neogene Sandakan Basin, South Sulu Sea, The Mem., 48, pp. 117–199.
Philippines. AAPG Search and Discovery Article, 90913. Thyberg, B., Jahren, J., winje, T., Bjorlykke, K., Inge, Faleide, J.I., Marcussen, O., 2010.
Samuel, O.J., Cornford, C., Jones, M., Adekeyec, O.A., Akandec, S.O., 2009. Improved Quartz cementation in Late Cretaceous mudstones, northern Norway: Changes in
understanding of the petroleum systems of the Niger Delta Basin, Nigeria. Org. rock properties due to dissolution of smectite and precipitation of micro-quartz
Geochem. 40, 461–483. crystals. Marine and Petroleum Geology 115, B10415. doi:10.1029/2009JB006997.
Sandal, S.T., 1996. The Geology and Hydrocarbon Resources of Negara Brunei Tingay, M.R.P., Hillis, R.R., Morley, C.K., Swarbrick, R.E., Drake, S.J., 2005. Present day
Darussalam, (1996 revision). Brunei Shell Petroleum Company/Brunei Museum, stress orientations in Brunei: A snapshot of prograding tectonics in a Tertiary delta.
Syabas Bandar Seri Begawan, Brunei Darussalam. 243 pp. J. Geol. Soc. Lond. 162, 39–49.
Sapiie, B., Adyagharini, A.C., Teas, P., 2010. New insight of tectonic evolution of Tingay, M., Hillis, R., Swarbrick, R., Morley, C.K., Damit, A., 2009. Origin of overpressure and
Cendrawasih Bay and its implication for hydrocarbon prospect, Papua, Indonesia. pore pressure prediction in the Baram Delta Province, Brunei. AAPG Bull. 93, 51–74.
Proceedings of the Petroleum Systems of SE Asia and Australasian Conference, May, Tingay, M.R., Morley, C.K., King, R.E., Hillis, R.R., Hall, R., Coblentz, D., 2010. The SE Asian
2010, Indonesia Petroleum Association, IPA10-G-158. 11 pp. Stress Map. Tectonophysics 482, 92–104.
Sapin, F., Pubellier, M., Ringenbach, J.-C., Bailly, V., 2009. Alternating thin versus thick- Tingay, M.R., Bentham, P., De Feyter, A., Kellner, A., 2010 (in press). Present-day stress
skinned decollements, example in a fast tectonic setting: The Misool-Onin- field rotations associated with evaporites in the offshore Nile delta. Geol. Soc. Am.
Kumawa Ridge (West Papua). J. Struct. Geol. 31, 444–459. Bull.
Schiefelbein, C.F., Zumberge, J.E., Cameron, N.C., Brown, S.W., 2000. Comparison of Totterdell, J.M., Krassay, A.A., 2003. The role of shale deformation and growth
crude oil along the south Atlantic margins. In: Mello, M.R., Katz, B.J. (Eds.), faulting in the Late Cretaceous evolution of the Bight Basin, offshore southern
Petroleum Systems of South Atlantic margins: AAPG Mem., 73, pp. 15–26. Australia. In: Van Rensbergen, P., Hillis, R.R., Maltman, A.J., Morley, C.K. (Eds.),
Screaton, E.J., Fisher, A.T., Carson, B., Becker, K., 1997. Barbados Ridge hydrogeological Subsurface Sediment Mobilization: Geological Society (London) Special Publica-
test implications for fluid migration along an active decollement. Geology 25, tion, 216, pp. 429–442.
239–242. Trudgill, B.D., Rowan, M.G., Fiduk, J.C., Weimer, P., Gale, P.E., Korn, B.E., Phair, R.L.,
Searle, M.P. and C. K. Morley, in press, Tectonic and thermal evolution of Thailand in the Gafford, W.T., Roberts, G.E., Dobbs, S.W., 1999. The Perdido Fold Belt, Northwestern
regional context of Southeast Asia. In: M.F. Ridd, A. J. Barber, M. J. Crow, (editors), Deep Gulf of Mexico. Part 1: Structural Geometry, Geolution and Regional
The Geology of Thailand, Geol. Soc. Lond. Mem. Implications: AAPG Bulletin, v. 83, pp. 88–113.
C.K. Morley et al. / Earth-Science Reviews 104 (2011) 41–91 91

Truempy, D.M., Mullin, P.R., Roberts, C., Ramsumair, R., 2004. A working petroleum Weber, J.C., Dixon, T.H., Demets, C., Ambeh, W.B., Mattioli, P.J.G., Saleh, J., Selle, G.,
system in deepwater East Trinidad (abstract). AAPG International Conference and Bilham, R., Perez, O., 2001. GPS estimate of relative motion between the Caribbean
Exhibition, Cancun, Mexico. and South American plates, and geological implications for Trinidad and Venezuela.
Underwood, M.B., 2002. Strike-parallel variations in clay minerals and fault vergence in Geology 29, 75–78.
the Cascadia subduction zone. Geology 30, 155–158. Weimer, P., Buffler, R., 1992. Structural geology and evolution of the Mississippi Fan
Underwood, M.B., 2007. Sediment inputs to subduction zones: why lithostratigraphy Fold Belt, deep Gulf of Mexico. AAPG Bull. 76, 225–251.
and clay mineralogy matters. In: Dixon, T.H., Moore, J.C. (Eds.), The seismogenic Weimer, P., Slatt, R.M., Bouroullec, R., Fillon, R., Pettingill, H., Pranter, M., Tari, G., 2006.
zone of subduction thrust faults. Columbia University Press, New York, pp. 42–85. Petroleum Geology of Deepwater Settings. AAPG Stud. Geol. 57 816 pp.
Van Hattum, M.W.A., Hall, R., Nichols, G.J., 2003. Provenance of northern Borneo White, R.S., 1982. Deformation of the Makran accretionary sediment prism in the Gulf
sediments. Proc. Indon. Petrol. Assoc., 29th Annual Convention and Exhibition, of Oman (north-west Indian Ocean). In: Legget, J.K. (Ed.), Trench and Fore-Arc
October 2003, IPA03-G-016, pp. 305–320. Geology: Sedimentation and Tectonics on Modern and Ancient Active Plate
Van Hattum, M.W.A., Hall, R., Pickard, A.L., Nichols, G.J., 2006. Southeast Asian Margins. Balckwell Scientific Publications, Oxford.
sediments not from Asia: Provenance and geochronology of north Borneo Wiener, R.W., Mann, M.G., Cotton, W.B., Booman, S.L., Lopez, C.J., Angelich, M.T.,
sandstones. Geology 34, 589–592. Molyneux, J., Barboza, S.A., 2009. Contractional domains of the Niger Delta:
Van Rensbergen, P., Morley, C.K., 2003. Styles and mechanisms of shale mobilization in structural styles, influence of mobile shale and structural-stratigraphic evolution.
the structural evolution of large deltas, a study of shale diapers in a delta setting, AAPG Hedberg Conference, Deepwater Fold and Thrust Belts, October 4-9, 2009,
offshore Brunei Darussalam. In: Van Rensbergen, P.R., Hillis, R., Maltman, A., Tirrenia, Italy, Abstracts.
Morley, C.K. (Eds.), Subsurface Sediment Mobilization: Geol. Soc. Lond. Spec. Pub., Wildi, W., 1983. Le chaine tello-rifaine (Algerie, Maroc, Tunisie): structure, strati-
216, pp. 395–409. graphie et evolution du Trias au Miocene. Rev. Geol. Dynam. Geog. Phys. 24,
Van Rensbergen, P., Morley, C.K., Ang, D.W., Hoan, T.Q., Lam, N.T., 1999. Structural 201–297.
evolution of shale diapirs from reactive rise to mud volcanism: 3D seismic data Wood, L.J., 2000. Chronostratigraphy and tectonostratigraphy of the Columbus Basin,
from the Baram Delta, offshore Brunei Darussalam. J. Geol. Soc. Lond. 156, eastern offshore Trinidad. AAPG Bull. 84, 1905–1928.
633–650. Wood, L.J., Mize-Spansky, K.L., 2009. Quantitative seismic geomorphology of a
Velaj, T., 2001. Evaporites in Albania and their impact on the thrusting process. J. Balkan Quaternary leveed-channel system, offshore eastern Trinidad and Tobago,
Geophys. Soc. 4, 9–18. northeastern South America. AAPG Bull. 93, 101–125.
Vendeville, B.C., 2005. Salt tectonics driven by sediment progradation: Part 1 – Worrall, D.M., Snelson, S., 1989. Evolution of the northern Gulf of Mexico, with emphasis
Mechanisms and kinematics. AAPG Bull. 89, 1071–1079. on Cenozoic growth faulting and the role of salt. In: Bally, A.W., Palmer, A.R. (Eds.), The
Vernant, P., Nilforoushan, F., Hatzfeld, D., Abbassi, M.R., Vigny, C., Masson, F., Nankali, H., Geology of North America: an overview: Geol. Soc. Am., The Geology of North America,
Martinod, J., Ashtiani, A., Bayser, R., Tavakoli, F., Chery, J., 2004. Present-day crustal A, pp. 97–138.
deformation and plate kinematics in the Middle East constrained by GPS Wu, S., Bally, A.W., Cramez, C., 1990. Allochthonous salt, structure and stratigraphy of
measurements in Iran and Northern Oman. Geophys. J. Int. 157, 381–398. the northeastern Gulf of Mexico, part II: structure. Mar. Petrol. Geol. 7, 334–370.
Vigny, C., Socquet, A., Rangin, C., Chamot-Rooke, N., Pubellier, M., Bouin, M.-N., Xie, X., Jiao, J.J., Li, S., Cheng, J., 2003. Salinity variation of formation water and
Bertrand, G., Becker, M., 2003. Present-day crustal deformation around Sagaing diagenesis reaction in abnormal pressure environments. Sci. China 46, 269–284.
Fault, Burma. J. Geophys. Res. 108. doi:10.1029/2002JB001999. Yue, L.-F., 2007. Active structural growth in central Taiwan in relationship to large
Von Blanckenburg, F., Davies, J.H., 1995. Slab breakoff; a model for syncollisional earthquakes and pore fluid pressures. Ph.D. Thesis, Princeton University, 273 pp.
magmatism and tectonics in the Alps. Tectonics 14, 120–131. Zahid, M.K., Uddin, A., 2005. Influence of overpressure on formation velocity evaluation
Wade, J.A., MacLean, B.C., 1990. The geology of the southeastern margin of Canada. In: of Neogene strata from the eastern Bengal Basin, Bangladesh. J. Asian Earth Sci. 25,
Keen, M.J., Williams, G.L. (Eds.), Geology of the continental margin of eastern 419–429.
Canada: Geol. Surv. Can., 2, pp. 167–238. Ziegler, P.A., 1989. Evolution of the North Atlantic – an Overview. In: Tankard, A.J.,
Warren, J.K., 2006. Evaporites: Sediments, Resources and Hydrocarbons. Springer. 1036 pp. Balkwill, H.R. (Eds.), Extensional Tectonics and Stratigraphy of the North Atlantic
Warren, J. K., Cheung, A., Cartwright, I., in press, Organic geochemical, isotopic and Margins: Am. Assoc. Petrol. Geol. Mem., 46, pp. 111–129.
seismic indicators of fluid flow in pressurized growth anticlines and mud volcanoes Ziegler, P.A., Roure, F., 1999. Petroleum systems of Alpine-Mediterranean foldbelts and
in modern deepwater slope and rise sediments of offshore Brunei Darussalam; basins. In: Durand, B., Jolivet, L., Horvàth, F., Séranne, M. (Eds.), The Mediterranean
Implications for hydrocarbon exploration in other mud and salt diapir provinces. Basins: Tertiary Extension within the Alpine Orogen: Geol. Soc., London, Spec-
AAPG Mem. Publ., 156, pp. 517–540.
Wawrzynice, T.F., Swenson, J., Ambrose, W., Fouad, K., Aranda, M., Zamora, E.M., Ziegler, P.A., van Wees, J.-D., Cloetingh, S., 1998. Mechanical controls on collision-
Sanchez-Barreda, L.A., 2003. Cenozoic deformation styles of the Laguna-Madre- related compressional intraplate deformation. Tectonophysics 300, 103–129.
Tuxpan shelf and Mexican Ridges fold belt, Mexico. Gulf Coast Assoc. Geol. Soc. Zielinski, G.W., Bjorøy, M., Zielinski, R.L.B., Ferriday, I.L., 2007. Heat flow and surface
Trans. 53, 843–854. hydrocarbons on the Brunei continental margin. AAPG Bull. 91, 1053–1080.

You might also like