Download as pdf or txt
Download as pdf or txt
You are on page 1of 27

Frontal expansion of an accretionary wedge under highly oblique

plate convergence: Southern Indo-Burman Ranges, Myanmar

Peng Zhang1,2,†, Shao-Yong Jiang1, Raymond A. Donelick3, Renyuan Li4, Cleber J. Soares5, and Lianfu Mei2
1State Key Laboratory of Geological Processes and Mineral Resources, Collaborative Innovation Center for Exploration of Strategic
Mineral Resources, China University of Geosciences, Wuhan 430074, China
2Key Laboratory of Tectonics and Petroleum Resources (China University of Geosciences), Ministry of Education,

Wuhan 430074, China


3Apatite.com Partners, LLC, 1075, Matson Road Viola, Idaho 83872-9709, USA
4International Limited, China National Offshore Oil Corporation (CNOOC), Beijing 10027, China
5ChronusCamp Research, São Paulo State 13974-160, Brazil

ABSTRACT successions and displays as a series of nega- sion is often interpreted as the cause of the
tive flower-like structures. These charac- steeply tapered wedges.
The formation of accretionary wedges teristic features are roughly consistent with
with oversteepened slopes and uplifted axial the results of laboratory analog modeling of INTRODUCTION
zones has been demonstrated to be poten- highly oblique plate convergence but signifi-
tially associated with highly oblique plate cantly differ from those of natural accretion- Convergent plate margins are depicted as
convergence by numerical and analog stud- ary wedges that formed under highly oblique principal regions of crustal growth and subduc-
ies. The direct role of this mechanism, or convergence conditions, such as those in Su- tion erosion and are central to our understand-
other factor(s) in producing the described matra, Hikurangi, Chile, and Cascadia. We ing of how plate tectonics operate on Earth over
structural and morphological features in further a­nalyzed the sediment provenance time (e.g., Frisch et al., 2011; Fitch, 1972; Stern,
nature, however, has yet to be confirmed. of the southern Indo-Burman Ranges and 2004; von Huene and Scholl, 1991). Wedge-
We used seismic reflection sections, detrital discovered that the outer-wedge rocks are a shaped sedimentary complexes, called accretion-
zircon U-Pb ages, and detrital apatite fission- product of sediment reworking of the hinter- ary wedges, often characterize such margins and
track thermochronological data to examine land wedge that began to be uplifted and ex- grow when materials are transferred from the
the effects of highly oblique convergence and humed in the early Miocene (22–12 Ma) due subducting plate to the overriding plate, either by
sediment reworking on accretionary wedge to transpressional motion between the Indian frontal offscraping at the trench axis or by under-
growth in the Indo-Burma Subduction Zone. plate and West Burma Terrane. Our analyses plating along the subduction décollement. The
A detailed subsurface structural analysis of indicate that active sedimentation behind the critical-taper wedge mechanics theory (Dahlen
a two-dimensional seismic survey from the major growth thrust fault (FT1) provided ad- et al., 1984; Davis et al., 1983) describes the
outer wedge of the southern Indo-Burman ditional basal shear stress that strengthened overall mechanics of wedges at convergent plate
Ranges, Myanmar, yielded three primary the coupling of the interface between the margins as analogous to a pile of soil/snow in
characteristics. These are (1) a narrow, steep wedge base and décollement and promoted front of a moving bulldozer. This theory predicts
deformation front (average width 15.6 km) the vertical expansion of the outer wedge that the critically tapered wedge will advance “in
and a vast, low-relief shelf terrace (average of the southern Indo-Burman Ranges from sequence” along the basal décollement, forming
width 49 km); (2) a comparatively long-lived the Neogene to the present day. In contrast, a broad seaward fold and thrust belt.
growth thrust fault (FT1) with a convex-up the outer wedge of the central Indo-Burman However, at highly oblique convergent mar-
geometry at the rear of the deformation front Ranges has experienced stronger forward ac- gins (convergence obliquity φ >60–70°; Braun
that controlled the vertical stack of the pro- cretion since the late Miocene, which could and Beaumont, 1995; Burbidge and Braun,
gradational sequences in the shelf terrace; be explained by a smaller degree of obliquity 1998), the relative motions of the lithospheric
and (3) a group of NE-striking transtensional and weaker sediment reworking. Our find- plates result in large margin-parallel strike-slip
faults that cut through entire outer-wedge ings demonstrate that both highly oblique faults at the rear of the wedge and prominently
plate convergence and sediment reworking attenuated margin-perpendicular crustal short-
were the primary driving forces that trig- ening (e.g., Fitch, 1972; McCaffrey, 1992).
Peng Zhang https://orcid.org/0000-0002- gered vertical development of accretionary Here, convergence obliquity, φ, represents the
7897-1144 wedges. The results of this research have sig- angle between the plate vector and the normal
†Present address: State Key Laboratory of Geological
nificant implications for understanding the to the plate boundary (e.g., Leever et al., 2011).
Processes and Mineral Resources, Collaborative
Innovation Center for Exploration of Strategic Mineral structures and kinematic evolution of wedge Accretionary wedges developed in this context
Resources, China University of Geosciences, Wuhan systems at other convergent plate margins, in are commonly doubly vergent and symmetric,
430074, China; p8.zhang@cug​.edu​.cn. which seamount passage or subduction ero- with oversteepened slopes and uplifted axial

GSA Bulletin; September/October 2023; v. 135; no. 9/10; p. 2348–2374; https://doi.org/10.1130/B36560.1; 14 figures; 2 tables; 1 supplemental file.
Published online 23 December 2022

For permission to copy, contact editing@geosociety.org


2348
© 2022 Geological Society of America

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Accretionary wedge development at oblique convergent margins

zones (Burbidge and Braun, 1998; Koons, 1994; Indo-Burman Ranges represent an accretionary of the hinterland wedge. Thus, we should: (1)
Leever et al., 2011; McClay et al., 2004; Mortera- wedge that developed under highly oblique plate verify that the Miocene rocks of the outer wedge
Gutiérrez et al., 2003). However, the direct role convergence since at least the Miocene (Bertrand have provenance features similar to those of the
of highly oblique plate convergence in producing and Rangin, 2003; Betka et al., 2018; Curray, Eocene rocks of the hinterland wedge (Target I),
the structures described has yet to be determined 2005; Maurin and Rangin, 2009a; Nielsen et al., and (2) confirm that the outer-wedge materials
in natural accretionary systems. Only a few stud- 2004) and potentially as early as the late Eocene were eroded from the hinterland wedge rather
ies have focused on the geometry and kinematic (ca. 40 Ma; Licht et al., 2019; Morley et al., than from other regions that have always been
evolution of accretionary wedges that developed 2020, 2021; Rangin, 2017; Westerweel et al., sources of the hinterland wedge (Target II). For
from highly oblique plate convergence, such as 2019). Although significant margin-parallel Target I, we used detrital zircon U-Pb ages from
those from Sumatra (Karig et al., 1980; McNeill components accommodated by strike-slip faults published samples and our new data set: if the
and Henstock, 2014; Moeremans et al., 2014), within the wedge itself and at the rear of the outer wedge showed age spectra similar to those
Hikurangi (Barnes et al., 2002; Barnes and de wedge have been determined (Betka et al., 2018; of the hinterland wedge, we could conclude that
Lépinay, 1997; Nicol et al., 2007), and Casca- Bürgi et al., 2021; Khin et al., 2020; Maurin and they share similar provenance features. For Tar-
dia (Adam et al., 2004; Booth-Rea et al., 2008), Rangin, 2009a; Morley et al., 2020; Morley, get II, we utilized detrital apatite low-tempera-
where the wedges narrow and steepen as the 2017; Nielsen et al., 2004; Pivnik et al., 1998; ture thermochronological data to demonstrate
convergence obliquity increases. Nevertheless, Vigny et al., 2003), crustal shortening in the that the hinterland wedge has uplifted since the
natural wedges at oblique convergent margins outer wedge of the central Indo-Burman Ranges early Miocene. If the hinterland wedge were
are not completely consistent with previous the- (20–25°N) dominated by a 120-km-wide fold uplifted in the Miocene, it would be able to sup-
oretical results (e.g., Leever et al., 2011; Mannu and thrust belt was still prominent. Some authors ply detritus to the outer wedge and, more impor-
et al., 2016; McClay et al., 2004; McNeill and have attributed this forward advance of the cen- tantly, stop detritus from other potential sources
Henstock, 2014; Morgan and Bangs, 2017). tral Indo-Burman Ranges to excess subsidence (e.g., West Burma Terrane) from being trans-
For example, wedges developed at the Sumatra and lithospheric sinking of the Surma subbasin ported into the outer wedge. Our results indicate
margin include broad deformation fronts and (Fig. 1; Mallick et al., 2020), an increase in the that highly oblique plate convergence and sedi-
prominent frontal accretion (e.g., Moeremans sedimentation rate in front of the outer wedge ment reworking have increased the duration and
et al., 2014). This discrepancy indicates that the (Najman et al., 2020), and/or a combined effect magnitude of slip on the thrust faults that previ-
structure and geometry of accretionary wedges of southwestward Tibetan Plateau crustal flow ously existed and delayed the formation of new
are affected not only by the convergence obliq- and strike-slip tectonics (Rangin et al., 2013; thrust faults in the deformation front, which led
uity but also by other factors, such as sediment Rangin and Sibuet, 2017). In the southern Indo- to the vertical growth of the outer wedge in the
reworking (i.e., active sedimentation and surface Burman Ranges (16–20°N), frontal accretion is southern Indo-Burman Ranges since the early
erosion). relatively limited, forming a narrow and steep Miocene (ca. 22 Ma). These findings may also
Numerical and analog models demonstrate outer wedge that may coincide with those at be used to understand wedge growth-associated
that active sedimentation can act as a first-order highly oblique convergent margins. However, sediment reworking at other highly oblique con-
control on wedge development. The exact mech- previous studies have usually attributed this vergent margins.
anism by which active sedimentation affects the wedge to the subduction of the Ninetyeast Ridge
wedge depends on the locus of deposition (Fillon beneath the Indo-Burman Ranges (Gopala Rao GEOLOGIC BACKGROUND
et al., 2013; Mannu et al., 2016; Mugnier et al., et al., 1997; Maurin and Rangin, 2009b). The
1997; Noda et al., 2020; Simpson, 2010; Storti kinematic development of the outer wedge as Tectonic Setting of the Indo-Burma
and McClay, 1995). Sedimentation on top of the a function of oblique plate convergence and/ Subduction Zone and Adjacent Regions
wedge would increase the taper to induce new or sediment reworking in both the central and
internal thrusts, promoting wedge deformation southern parts of the Indo-Burman Ranges has The Indo-Burman Ranges and West Burma
over the décollement, whereas sedimentation in not been described in detail in previous studies. Terrane constitute the Indo-Burma Subduction
the trench would lead to deformation concen- In this study, we combined field mapping, Zone, which connects the Eastern Himalayan
trated in front of the wedge and enhance forward two-dimensional (2-D) seismic-reflection sec- Syntaxis and Sunda trench-arc system to the
advance of the wedge (Fillon et al., 2013; Mor- tions, detrital zircon U-Pb ages, and detrital apa- north and south, respectively. The Himalayas
ley, 2007; Noda et al., 2020; Simpson, 2010). tite low-temperature thermochronological data were formed in the Cenozoic after the India-
If the wedge continues its vertical deformation to test the effects of highly oblique plate con- Eurasia collision and include three distinct tec-
above sea level, surface erosion occurs in the vergence and sediment reworking of the hinter- tonostratigraphic units (Fig. 1). In the north, the
hinterland and restrains the wedge from attain- land wedge rocks on outer-wedge expansion in Xigaze forearc basin and Indus-Yarlung suture
ing critical equilibrium (Cruz et al., 2010, 2011; the southern Indo-Burman Ranges. While field zone separate the Himalayas from the Gang-
Davis et al., 1983; Mugnier et al., 1997). How- investigations and 2-D seismic data can be used dese Arc. The Assam and Bengal basins south
ever, the coupling effects of active sedimenta- to characterize highly oblique plate convergence of the eastern Himalayas comprise ∼10–20 km
tion and surface erosion on accretionary wedges by identifying flower-like structures and spe- of shallow marine to deltaic-fluvial sedimentary
remain unresolved. cific fault assemblages, the coupled application rocks. The southern end of the Bengal Basin
Here, we investigate the deformation history of detrital zircon U-Pb ages and detrital apatite empties into the Bengal Fan, and farther south,
of the Indo-Burman Ranges (Fig. 1), which are fission-track results can depict surface erosion into the Nicobar Fan (Fig. 1), all of which were
key to understanding the eastern Tethys sub- and redistribution of the hinterland wedge mate- mainly formed during the Neogene to Quater-
duction and subsequent India-Eurasia collision rials. To test the assumed sediment reworking in nary (Blum et al., 2018; McNeill et al., 2017;
(Advokaat et al., 2018; Morley et al., 2020, 2021; the southern Indo-Burman Ranges, the key is to Pickering et al., 2020).
Najman et al., 2022; Westerweel et al., 2019). demonstrate that the Miocene rocks of the outer The Sunda trench-arc system in the Suma-
Several researchers have recognized that the wedge were recycled from the Eocene rocks tra region includes the Woyla Arc and West

Geological Society of America Bulletin, v. 135, no. 9/10 2349

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Peng Zhang et al.

Figure 1. Geological map of


Indo-Burma Subduction Zone
and the eastern Himalayan
Syntaxis shows major terranes,
terrane boundaries, geologi-
cal units, and rivers (modified
after Advokaat et al., 2018;
Curray, 2005; and Zhang et al.,
2021a). Seismic profiles and
drillings used in Figure 5 are
marked. Sediment isopaches
presented in the Bay of Ben-
gal and Andaman Sea are af-
ter Curray (1991) and Morley
and Alvey (2015). Inset map
shows plate motion vectors
of Myanmar and India in the
Sundaland reference frame (af-
ter Vigny et al., 2003). KGR—
Katha Gangaw Range; N.
MMB—northern Mogok Meta-
morphic Belt within SE Tibet;
S. MMB—southern Mogok
Metamorphic Belt in the Shan
Scarps; WPA—Wuntho-Popa
Arc; TMB—Tagaung-Myitky-
ina Belt; KR—Kumon Range;
JB—Jade Belt; MR—Minwun
Range; WR—Wuntho Range;
EHS—Eastern Himalayan Syn-
taxis; THS—Tethyan Himala-
yan Sequence; GHC—Greater
Himalayan Crystalline Com-
plex; LHS—Lesser Himalayan
Sequence; SHS—Sub-Himala-
yan Sequences; GCT—Great
Counter Thrust; STD—South
Tibet Detachment; MCT—
Main Central Thrust; MBT—
Main Boundary Thrust;
MFT—Main Frontal Thrust;
HK—Hukawng; CH—Chind-
win; SM—Shwebo-Minwun;
MB—Minbu; BY—Bago-
Yoma; PI—Prome-Irrawaddy.
Seismic profiles A–H are illus-
trated in Figure 5.

2350 Geological Society of America Bulletin, v. 135, no. 9/10

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Accretionary wedge development at oblique convergent margins

­ umatra, which assembled in the mid-Creta-


S h­ interland wedge in the southern Indo-Burman shore of Myanmar (outer wedge), tectonostrati-
ceous (Fig. 1; Advokaat et al., 2018; Barber and Ranges is unclear; thick (3–5 m) sandstone beds graphic sequences include units separated by
Crow, 2009). The Woyla Arc represents an intra- and interbedded mudstone beds with blocks of unconformities of regional extent (or their cor-
oceanic volcanic arc within the eastern Tethys serpentinite have been reported on its western relative conformities; horizons U1 to U6) that
(Barber and Crow, 2009). Igneous rocks exposed margin (Moore et al., 2019; Naing, 2019; Zhang record deposition during different stages in the
in the West Sumatra and Woyla Arc yielded pre- et al., 2021a). evolution of the accretionary wedge (Fig. 2B).
dominant magmatism in the Late Triassic, mid- The outer wedge is separated from the hinter- These six horizons divide the outer-wedge suc-
Cretaceous, and Cenozoic (Zhang et al., 2019a). land wedge by the Kaladan Fault (Fig. 1). The cession into seven packages with distinct ampli-
These rocks suggest that the Woyla Arc formed outer wedge was best investigated in the Chit- tudes, reflector continuities, and deformation
above a SW-dipping subduction zone in modern tagong-Tripura Belt (Bangladesh and NE India). characteristics.
coordinates, and West Sumatra above a synchro- East-dipping thrusts converging to a décollement Along the Pyay–Toungup section (Fig. 3),
nous NE-dipping subduction zone (Advokaat in Oligocene shales dominate this belt (Betka the majority of the sediments (Rakhine ­Flysch)
et al., 2018). The Sunda margins have absorbed et al., 2018; Maurin and Rangin, 2009a). Off- were mapped from the Late Cretaceous to
the oblique convergence between the Indian- shore, in western Myanmar, however, the outer Eocene (Bannert et al., 2011; Brunnschweiler,
Australian plate and Sundaland (Fig. 1) since the wedge is characterized by gentle folds and grav- 1966; Burma Earth Sciences Research Divi-
early Neogene by partitioning over the forearc ity sliding (Maurin and Rangin, 2009b; Morley sion, 1977; United Nations, 1979) based on the
areas and several right-lateral faults (such as et al., 2020; Zhang et al., 2021a). identification of Globotruncana-Nummulitic–
the Sumatra and West Andaman faults; Misawa The West Burma Terrane, separated by the bearing limestones and microfossil-bearing
et al., 2014; Vigny et al., 2003). west-dipping Kabaw Fault from the Indo- shales (Ammodiscus siliceous, Globotruncana
The Indo-Burman Ranges consist primarily Burman Ranges, is inferred to be either a 2, Gumbelina 7, Heterohelix globulosa, Pre-
of an Eastern Belt (wedge core), Western Belt continental block related to Cathaysia, which discosphaera columnata, and Watznaueria sp.;
(hinterland wedge) to the west, and farthest formed by the Late Triassic (Barber and Crow, Bannert et al., 2011; Gramann, 1974; Naing,
west the Outer Belt (outer wedge; Maurin and 2009), or a part of the Trans-Tethyan Arc that 2019; Fig. S2; see footnote 1). However, the
Rangin, 2009a; Mitchell, 2018). In the central rifted from the northern margin of Gondwana assigned ages are inconsistent with the maxi-
and northern Indo-Burman Ranges (Fig. 1), the before the mid-Cretaceous (Licht et al., 2020; mum depositional ages (MDAs) determined
wedge core itself is composed of the Upper Tri- Westerweel et al., 2019; Zhang et al., 2021b). from detrital zircon U-Pb data (49.6–42.5 Ma;
assic Kanpetlet Schist and Pane Chaung turbi- The West Burma Terrane rocks exposed in the Naing, 2019; Najman et al., 2020; this study).
dites, upper Albian Paung Chaung limestones, Wuntho Range are undated Hypu Taung and This discrepancy could be explained by either
and minor ophiolites within or overlying the Shwedaung metamorphic rocks, Jurassic to unmapped structural interleaving of Eocene
turbidites and limestones (Mitchell, 2018; Mor- mid-Cretaceous Mawgyi and Mawlin volca- and Cretaceous rocks (Najman et al., 2020)
ley et al., 2020). However, the origins of and nogenic units, Upper Cretaceous Namakauk or the large age ranges of the microfossils
relationships among the Mesozoic rocks are limestone, and Upper Cretaceous to lower (Naing, 2019).
disputed, and the Paung Chaung limestones Paleogene clastic formations. Their equiva- In the Rakhine coast region of the southern
and equivalents are unconformably underlain lents or parts of these units are reported in Indo-Burman Ranges, the Eocene rocks (Nga-
by the Kanpetlet Schist and the Pane Chaung the Jade Belt, Kumon Range, Minwun Range, pali Formation) were drilled by well RM-1 and
turbidites in most places (Mitchell, 2018; Mor- Katha-Gangaw Range, and Tagaung-Myitky- showed paleontological assemblages of Globi-
ley et al., 2020). The Pyay–Toungup section of ina Belt (Fig. 1). The West Burma Terrane is grtinatheka subconglobata curry, Turborota-
the southern Indo-Burman Ranges investigated currently occupied by the Central Myanmar lia cerroazulensis cerroazulensis, Globigerina
in this study lacks the typical wedge core rocks Basin. This Cenozoic forearc-backarc basin inaequispira, and Turborotalia cerroazulensis
(Figs. 2–4 and Supplementary Material Fig. S11). consists of several isolated subbasins (Fig. 1; frontosa (Lei et al., 2009; Fig. S2), which indi-
The hinterland wedge is separated from the Bertrand and Rangin, 2003; Pivnik et al., 1998; cate a Lutetian–Bartonian age. The inferred
wedge core by the Kheng Fault and its equiva- Zhang et al., 2021a). Eocene age is consistent with the MDAs deter-
lents (Fig. 1). It contains a complex of folded mined from detrital zircon U-Pb analyses (47.2–
and thrusted Upper Cretaceous to Paleogene Stratigraphy and Stratigraphic Age 41.2 Ma; Najman et al., 2020; this study). The
turbidites, with numerous blocks of limestone, Control of the Southern Indo-Burman lower Oligocene in this region was previously
andesite, pillow-basalt, chert, and serpentinite Ranges thought to have been absent, and the Eocene
(Mitchell, 2018; Moore et al., 2019). The Falam was unconformably overlain by the upper Oli-
and Chunsung formations represent these turbi- A fully integrated stratigraphy has not been gocene Yechangyi Formation (Yang et al., 2009).
dites in the central Indo-Burman Ranges (Mitch- established for the hinterland wedge of the However, paleontological assemblages of Glo-
ell, 2018). The stratigraphic division of the southern Indo-Burman Ranges, and much of bigerina gortanii, Globorotalia opima nana,
the nomenclature described here (Fig. 2A) and Globigerina ciperoensis ciperoensis (Lei
1Supplemental Material. Descriptions of field
is the result of published data (Naing, 2019; et al., 2009) suggest that sedimentation occurred
work and analytical methods; additional figures Naing et al., 2014) and several internal reports from the Rupelian to the Chattian. Up section,
showing cross sections, 2D seismic sections, and provided by the China National Offshore Oil the Miocene rocks (Yenandaung, Maragyun,
multidimensional scaling map; and tables containing Corporation (Qiu et al., 2008; Yang et al., 2007, and Leikkamaw formations; Fig. 2A) comprise
full results of new detrital zircon U-Pb analyses, apatite 2009; Zhao et al., 2006; Zheng, 2008). We used paleontological assemblages of Globigerinoides
fission track results, and compiled published detrital
zircon U-Pb ages. Please visit https://doi​.org​/10​ these data and (un)published geological maps primordius, Globigerina euapertura, Globige-
.1130​/GSAB​.S.21326094 to access the supplemental (Burma Earth Sciences Research Division, rina praebulloides praebulloides, Globigerina
material, and contact editing@geosociety​.org with 1977; Myint Swe, 2007; Tsung Aung, 1982) to praebulloides occulusa, Globorotalia mayeri,
any questions. create an updated geological map (Fig. 3). Off- Globorotalia obesa, Discoaster surculus, B ­ revis,

Geological Society of America Bulletin, v. 135, no. 9/10 2351

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Peng Zhang et al.

Figure 2. (A) Tectonostrati-


graphic chart for hinterland
wedge of the southern Indo-
Burman Ranges (IBR) was
compiled from Naing et al.
(2014) and Zhang et al. (2019b).
(B) Seismostratigraphy of the
outer wedge of the southern
B Indo-Burman Ranges. Six
main unconformities (U1, U2,
U3, U4, U5, and U6) were in-
terpreted across the study area,
and these unconformities rep-
resent the top Eocene, top Oli-
gocene, top lower Miocene, top
upper Miocene, top lower Plio-
cene, and top upper Pliocene,
respectively. TWT—two-way
traveltime; MDA—maximum
depositional age.

Florschuetzia spp., and Sphenolithus tintin- cal data or assigned by (un)published geologi- Offshore of western Myanmar (outer wedge),
abulum (Lei et al., 2009; Najman et al., 2020; cal maps. We do not consider the depositional the stratigraphic divisions and age controls
Figs. 2A and S2), all of which match the ages ages inferred from detrital zircon U-Pb data to be were primarily derived from industrial drilling
assigned in the Burma Earth Sciences Research credible given the presence of the older zircons well A4-H1 and our new 2-D seismic profiles
Division geological map. Depositional ages associated with sediment reworking in the post- (Figs. 2B and S3; see footnote 1). Regional
inferred from our detrital zircon U-Pb analyses Eocene rocks (Najman et al., 2020; this study; unconformities that separate the different
are older than those obtained from paleontologi- see below). seismic reflection layers act as stratigraphic

2352 Geological Society of America Bulletin, v. 135, no. 9/10

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Accretionary wedge development at oblique convergent margins

Figure 3. Detailed geological map of the southern Indo-Burman Ranges (after Burma Earth Sciences Research Division, 1977; Myint Swe,
2007; Tsung Aung, 1982; and Najman et al., 2020) is presented with seismic reflection data used in Figures 6–8 and detrital samples used
in Figures 9 and 12. Inset map shows stereograms (Wulf projections, lower hemisphere) of fault-slip vectors measured at five groups of
conjugate faults, and σ1, σ2, and σ3, represent maximum, intermediate, and minimum stress, respectively.

­boundaries. Asterorotalis trispinosa and ben- road-cut outcrops (Fig. 4) along the transects Two-Dimensional Reflection Seismic Data
thonic forms (Bolivina, Lagena, Nonion, Cibi- that crossed each primary structure. Lithologi-
cides, and Trifarina) were found in the Pleis- cal contacts were not extrapolated because rocks Two-dimensional seismic sections were
tocene rocks and Pulleniatina obliquiloculata, in the hinterland wedge were strongly deformed acquired during many petroleum exploration
Globorotalia tumida, and Globigerinoides were and commonly included cut through thrusts and campaigns over the past four decades. All reflec-
discovered in the Pliocene rocks (Yang et al., strike-slip faults. Eight cross sections (P1–P8) tion profiles were processed by deconvolution,
2009). Sedimentary rocks below horizon U4 were constructed perpendicular to the structural multiple suppression, time migration, and filter-
(Eocene to Miocene) were not fully drilled by axes (Fig. S1). Structural geology software 2-D ing to improve them to the highest quality (e.g.,
well A4-H1; therefore, the stratigraphic divi- Move and manual cross-section construction Zhang et al., 2021a). Seismostratigraphic analy-
sions were primarily correlated with published were used to interpret the kinematic evolution, sis was conducted using GeoFrame_4.5 soft-
seismic sections from adjacent regions (Maurin tectonic shortening, and surface erosion for each ware, which allows identification of stratigraphic
and Rangin, 2009a; Nyan Tun et al., 2007; Yang anticline and syncline. Descriptions of the anti- boundaries that show onlap relationships, trun-
and Kim, 2014), as well as with the surface geol- clines and synclines are summarized in detail in cated reflectors, and contrasts between seismic
ogy of the Ramree and Cheduba Islands. Text S1 (see footnote 1). We collected fault-slip facies reflectivity. Surface geology (Rakhine
data (slip parameters included dip directions coastal area; Fig. 3) and formation-top data from
MATERIALS AND ANALYTICAL and angles of the faults; n = 28) and the sense drilled wells (RM-1, RM-3, H-1, and adjacent
METHODS of motion inferred from kinematic indicators wells) were used to constrain the main bound-
(overturned folds, Riedel shears, offset markers, aries (Fig. 5). Sediments deposited in the outer
Geologic Mapping and Analysis of Tectonic and fibrous minerals) to characterize the tectonic wedge can be divided into seven units (Eocene,
Stress Field stress field during sedimentation and faulting Oligocene, lower–middle Miocene, upper Mio-
(Fig. 4). The computer program TectonicsFP cene, lower Pliocene, upper Pliocene, and Qua-
A new geological map (Fig. 3) was used to was employed to calculate the directions of the ternary; Fig. 2B). We compiled eight regional
describe the hinterland wedge in the study area. three principal stress axes for five groups of con- geological cross sections across the westernmost
Bedding dips (n = 80) were obtained from jugated shear joints (Fig. 3). part of the hinterland wedge, outer wedge, and

Geological Society of America Bulletin, v. 135, no. 9/10 2353

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Peng Zhang et al.

1m SE NW NW 325°
Ax1 Thrust W
Ax2 a a

a′

45°
°
120 a′
b′ b
σ1 L σ1 E
62
° b′
b

15

A B C
Overturned anticline SE
E NE
SW

D E F G
E W SE

H I J

Figure 4. Field photographs show lithofacies and structures from the hinterland wedge; see Figure 3 for locations. (A) Fold and thrust
within Eocene strata of the hinterland wedge, west of Pyay. (B) Two groups of conjugate faults in the Eocene strata; the older group
(black dotted lines) was formed by approximate north–south contraction, and the younger group (white dashed lines) was formed by
approximate east–west contraction. (C, E) Conjugate faults in the Eocene strata. (D) Overturned anticline and associated thrust. (F)
Antistep dextral translational fault with glided calcite crystal. (G, H) Thrust faults with numerous cleavages. (I, J) Conjugate faults
within the Miocene strata.

abyssal plain (Bengal Fan) to illustrate the gen- seismic data is generally good, particularly from western Myanmar and adjacent regions (e.g.,
eral structural geometry and kinematic evolution the seafloor down to the top Miocene unconfor- Blum et al., 2018; Cina et al., 2009; Lang and
of the southern Indo-Burman Ranges (16–20°N; mity (U4 horizon) at ∼2.0–2.5 sTWT. Below Huntington, 2014; McNeill et al., 2017; Morley
Fig. 5). We then concentrated on the outer wedge this level, the data quality commonly deterio- and Arboit, 2019; Najman et al., 2020; Pick-
in the Rakhine Block A4 (RBA4) between 18°10′ rates, although reliable recognition of the major ering et al., 2020; Zhang et al., 2019b). We
N and 19°00′ N (Fig. 3), where newly acquired faults is still possible down to 4 sTWT (Fig. 2B). collected 10 samples for detrital zircon U-Pb
2-D reflection seismic profiles covering an area analyses (Table S1; see footnote 1) to constrain
of 20 × 100 km were available, to describe its Detrital Zircon U-Pb Geochronology their derivations and assess sediment rework-
structures, fault activity, and sedimentation ing of the hinterland wedge rocks in the south-
(Figs. 6–8). This survey included 57 NE–SW- Detrital zircon U-Pb ages record the time of ern Indo-Burman Ranges. Clastic rocks were
striking cross-lines and 21 NW–SE-striking in- zircon growth in magma, or less commonly, of collected along the Pyay–Toungup–Kyaukpyu
lines, both spaced at 2 km and extending to 6 s zircon growth during metamorphism, and are Road (Fig. 3) because: (1) previous studies
two-way travel time (TWT). The quality of the useful for assessing sedimentary provenance in have provided basic geological information

2354 Geological Society of America Bulletin, v. 135, no. 9/10

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Accretionary wedge development at oblique convergent margins

A0 Deformation front
sea level
Gentle shelf B0 AP Deformation front Gentle shelf Outer fold-thrust belt
sea level
1 Shwe-4A 1
TWT [s]

TWT[s]
2
3 3
4 4
5 5

C0 AP DF Gentle shelf
RM3-1 Ramree
Island E0 sea level
AP Deformation front Gentle shelf H-1
sea level
1 1
Shale
2 Shale
2 diapir
TWT [s]

TWT [s]
diapir
3 K 3
al Shale
ad
4 an 4 diapir
FT3 FT2 FT1 5
5 FT3 FT2 FT1
De tac hm ent lev
6 el 6 Detachment level

D 0Deformation front Gentle shelf


RM1-1 Circular syncline
F0
on Ramree Island
sea level Deformation front Gentle shelf
1 sea level Rubis-1(projected)
1
2 Shale
TWT [s]

TWT [s]
2
3 diapir
Shale 3
4 diapirs
Ka 4
5 FT2 FT1 lad
FT3 an 5
Detachm ent level FT2 FT1 Detachment level
6 FT3
Deformation
G0 Abyssal plain front Gentle shelf H0 DF Gentle shelf
A7-1
sea level sea level
1 1
2 2
TWT [s]
TWT [s]

3 3
4 4
5 5
6 FT1 6 FT1
FT3 FT2 FT3&FT2 Detachment level
7 Detachment level 7
10 km Eocene Oligocene Lower-Middle Miocene Upper Miocene Lower Pliocene Upper Pliocene Pleistocene Holocene Industrial well

Figure 5. Line drawing of regional seismic sections shows geometry, structure, and Eocene–Quaternary successions of the outer wedge
of the southern Indo-Burman Ranges. The interpreted thrust faults within the deformation front are numbered in order (FT1, FT2, and
FT3). Seismic sequences were correlated with offshore Myanmar industrial wells and the surface geology of Ramree and Cheduba islands.
Profiles A and B were modified after Yang and Kim (2014) and Jain et al. (2010), respectively, and others were provided by CNOOC; all
uninterpreted seismic profiles are shown in the Supplementary Material (see footnote 1). See Figure 1 for seismic locations. TWT—two-way
traveltime; AP—abyssal plain; DF—deformation front; RM—Ramree Island.

on tectonic deformation and stratigraphy of to exclude grains with a >10% discordance. ing and rates of rock cooling through the AFT
this region (e.g., Brunnschweiler, 1966; Naing, This yielded at least 87 grains per sample, partial annealing zone, which ranges between
2019); (2) Eocene to Miocene rocks have been resulting in 95% confidence that any age popu- 60 °C and 120 °C (e.g., Reiners and Brandon,
roughly distinguished and mapped along this lations comprising more than 7% of the sam- 2006). To constrain exhumation history of the
transect (e.g., Burma Earth Sciences Research ple will be measured (Vermeesch, 2004). The hinterland wedge of the southern Indo-Burman
Division, 1977; United Nations, 1979), 206Pb/238U ratio was used to calculate the age Ranges and test the hypothesis of sediment
although there are still a lot of uncertainties; of grains younger than 1000 Ma, whereas the reworking since the early Miocene, 11 apatite
and (3) surface geology of this region could 207Pb/206Pb ratio was used for older grains. The samples were prepared and analyzed using the
be compared with newly acquired 2-D seismic new data set and previously published samples LA-ICP-MS method at either Apatite.com Part-
sections and drilling data in the RBA4 area (excluding grains with a >10% discordance) ners (USA) or ChronousCamp Research (Bra-
(offshore), which provides a unique chance to were processed and visualized using the R zil). We refer the reader to the Supplemental
validate the hypothetic scenario involving sedi- package IsoplotR (Vermeesch, 2018; Figs. 9– Material for the detailed method and full results
ment reworking (hinterland-wedge erosion and 11). The weighted average 206Pb/238U age for (Text S3 and Table S3; see footnote 1). Specifi-
outer-wedge sedimentation). the youngest group of at least two zircon grains cally, of the 11 analyzed samples, seven were
Laser ablation–inductively coupled plasma– overlapping within error at the 1 sigma level collected from the hinterland wedge (M02, M07,
mass spectrometry (LA-ICP-MS) was used to (YC1σ[2+]; Dickinson and Gehrels, 2009) is M11, M23, M25, M28, and M211), two from
measure zircon U-Pb ages at the State Key quoted here as the MDA. the outer wedge (Ramree Island; M08 and M12),
Laboratory of Geological Processes and Min- and the remaining two from the forearc subba-
eral Resources, China University of Geosci- Detrital Apatite Fission-Track sin (southern Minbu; M220 and M221). As sam-
ences, Wuhan. A description of the method is Thermochronology pling in the deformation front of the outer wedge
given in Text S2 (see footnote 1), and the full (offshore area) is unavailable, the deformation
results are listed in Table S2 (see footnote 1). Apatite fission-track (AFT) thermochronol- history of this region was constrained mainly by
The data described in this article were filtered ogy provides invaluable constraints on the tim- our 2-D seismic reflection data.

Geological Society of America Bulletin, v. 135, no. 9/10 2355

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Peng Zhang et al.

A Figure 6. (A) Short seismic reflection pro-


file across the deformation front and Abys-
sal plain is shown. The deformation front is
separated from the unfolded abyssal plain
by a landward-dipping thrust, which shows
upward propagation but has yet to rupture
seafloor sediments. The thick abyssal plain
sediment section (>6.0 km) near the defor-
mation front appears to be characterized by
numerous cross-cutting faults. The abyssal
plain sediment section that has been incor-
porated into the deformation front between
FT2 and FT3 is ∼5 km in width, which sug-
gests a very limited frontal accretion in the
Neogene. (B) Expanded view of deformation
front, behind which fanning growth pack-
ages of sediment deposited within the thrust-
top basin are observed. Prominent onlap
reflectors along main boundaries and pro-
gradation reflections within several seismic
B
units such as the Oligocene succession (inset
map) are identified. Erosional surfaces with
distinct truncation reflections and a buried
channel are also interpreted. (C) Short seis-
mic reflection profile across the rear part of
the shelf domain shows large-scale listric
normal faults that appear to have converged
at the deeper décollement surface. These
normal faults cut through the Oligocene–
lower Pliocene sequences, which indicates a
distinct block rotation. Overlying the listric
normal faults, large-scale propagation re-
flectors and erosion surfaces are observed.
See Figure 3 for locations of the seismic
­profiles. CDP—common depth point.

Detrital apatite grains were sorted into dif-


ferent groups using the following steps: (1)
measure rare earth element concentrations
C and 206Pb*/238U ages for each of ∼9 differ-
ent apatite standard species analyzed during
a LA-ICP-MS session, and then calculate the
mean and standard deviation for each of the
measured standard value; (2) derive a session-
specific, best-fit line for each measurement that
gives standard deviation as a function of mean
among all standards; (3) for each unknown
apatite grain from our new sample, use the
derived best-fit lines to define an apatite grain
group centered at the measured values for the
unknown grain and bounded by two standard
deviations about the measured values; and (4)
search for and isolate other unknown apatite
grains that exhibit measured values that fall
within the bounds of the defined apatite grain
group. Using the color ramp for 206Pb*/238U
age, rare earth element plots for all unknown

2356 Geological Society of America Bulletin, v. 135, no. 9/10

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Accretionary wedge development at oblique convergent margins

Figure 7. NW–SE-striking seismic profiles across the outer wedge of the southern Indo-Burman Ranges show planar faults that cut
through the entire wedge successions, with some of them forming negative flower-like structures. Major transtensional faults dis-
cussed in Figure 8 are numbered (FS1, FS2, FS3, and FS4). See Figure 3 for locations of the seismic profiles. Panel A shows planer
faults formed at the rear of the deformation front, whereas panel B shows planer faults developed close to the trench axis. CDP—
Common depth point.

apatites are shown in the Supplementary Mate- of the southern Indo-Burman Ranges. FT3 is ventional AFT and high resolution-AFT data.
rial. In this article, FT3 was used to describe Apatite.com Partners’ software package for The philosophy of this software is similar to
the degree of burial reheating and exhumation calculating thermal histories based on con- that of AFTSolve (Ketcham et al., 2000) but

Geological Society of America Bulletin, v. 135, no. 9/10 2357

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Peng Zhang et al.

A B

C D

Figure 8. (A–C) Time–structure maps show top upper Miocene, top lower Pliocene, and top upper Pliocene, respectively, of the RBA4 area
from 2-D seismic interpretation. (D) Simple tectonic model of the RBA4 area during the early Pliocene shows five blocks cut by the trans-
tensional faults (FS1, FS2, FS3, and FS4), consistent with the Riedel shears in a right-lateral shear system. The early Pliocene slope-break
line and transtensional faults indicate that these blocks are moving in distinct directions. TR—Thrust faults.

allows provenance history to vary among mod- AFT age goodness-of-fit (GOF) and confined GOF values are ≥0.05 in this study. The best fit
eled apatite grain populations. FT3 yields time- fission-track length GOF values for each popu- model is the model yielding the greatest prod-
temperature (t-T) paths with corresponding lation. Models are deemed acceptable when all uct of all GOF values.

2358 Geological Society of America Bulletin, v. 135, no. 9/10

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Accretionary wedge development at oblique convergent margins

B C

Figure 9. Kernel density estimates (KDE) of zircon U-Pb data show the new samples we analyzed (marked with asterisks) and
published samples (Naing et al., 2014; Najman et al., 2020; Zhang et al., 2018, 2019b) in (A) the southern Indo-Burman Ranges
(IBR) and (B) southern Minbu subbasin. Insets display detailed U-Pb age spectra in the range of 0–300 Ma. N—total analyzed
zircons; n—zircons younger than 300 Ma; n′—zircons calculated for maximum depositional age (MDA). The weighted average
206Pb/238U age for the youngest group of at least two zircon grains overlapping within error at the 1 sigma level (YC1σ[2+];

Dickinson and Gehrels, 2009) is quoted here as the MDA. (C) Same data as in panels A and B is displayed as a multidimensional
scaling (MDS) map calculated using the Kolmogorov–Smirnov statistic (Vermeesch, 2013). An MDS plot visualizes the degree
of similarity between samples, with the proximity of sample points reflecting their similarity. The axis scales are dimensionless
and have no physical meaning. Stress = 10.038287; a goodness-of-fit or stress of <10% is considered “good” for moderate to
large data sets (Vermeesch, 2013). The KDE plots were made using IsoplotR (Vermeesch, 2018).

Geological Society of America Bulletin, v. 135, no. 9/10 2359

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Peng Zhang et al.

A B C

Figure 10. Zircon kernel density estimate (KDE) plots compare our new data sets with published data in the (A)
Himalayan foreland (Bracciali et al., 2015; Cina et al., 2009; Lang and Huntington, 2014; and Najman et al.,
2008), submarine fans (Blum et al., 2018; McNeill et al., 2017; Pickering et al., 2020), and Andaman Flysch (Li-
monta et al., 2017; Bandopadhyay et al., 2022); (B) Indo-Burman Ranges (Aitchison et al., 2019; Allen et al., 2008;
Lin et al., 2022; Najman et al., 2020; Rahman et al., 2020; Vadlamani et al., 2015; Yang et al., 2020); (C) Central
Myanmar Basin (Arboit et al., 2021; Cai et al., 2020; Licht et al., 2019; Robinson et al., 2014; Wang et al., 2014;
Westerweel et al., 2020; Zhang et al., 2019b); and (D) from modern river sands (Bao et al., 2015; Bracciali et al.,
2015; Cina et al., 2009; Garzanti et al., 2016; Lang and Huntington, 2014; Lin et al., 2017; Zhang et al., 2012). (E)
The same data in panels A–D are displayed as a multidimensional scaling (MDS) map. The MDS map indicates
that the southern Indo-Burman Ranges and Central Myanmar Basin rocks were mainly derived from the Burma
Terrane. Details of these published data are given in Table S4 (see footnote 1). Fm—formation.

2360 Geological Society of America Bulletin, v. 135, no. 9/10

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Accretionary wedge development at oblique convergent margins

RESULTS

Structural Analysis of the Southern


Indo-Burman Ranges

Hinterland Wedge, Onshore Southern


Indo-Burman Ranges
The hinterland wedge of the southern Indo-
Burman Ranges consists of highly deformed
NW-SE–striking anticlines (sections P1–P6; Figure 11. Zircon kernel den-
Fig. S1). These anticlines are asymmetric, with sity estimate (KDE) plots for
shorter forelimbs and longer back limbs that are potential source regions are
both, in places (sections P1–P3), dominated by shown. Published data: Lhasa
second-order synclines or overturned synclines. Terrane (Fan et al., 2017; Geh-
All antiforms are separated by east-dipping rels et al., 2011; Leier et al.,
thrusts with fault planes that dip as much as 90° 2007; Wang et al., 2016); Gang-
(45–55° on average) and form a series of west- dese Arc and Dianxi-Burma
vergent, fault-related folds (Text S1 and Fig. Batholith (see references in
S1). Gentle synclines extend through the west- Zhang et al., 2019b); Himalaya
ernmost part of the hinterland wedge and are (Cai et al., 2016; Gehrels et al.,
cored by Paleogene sediments (Fig. S1). Total 2011; McQuarrie et al., 2008,
horizontal displacements on these major thrusts 2013; Myrow et al., 2010; Wang
are estimated to be ∼19.7 km by restoring hang- et al., 2016; Webb et al., 2013);
ing wall/footwall to the initial position, whereas Indian Craton (Myrow et al.,
the amount of fold shortening is estimated to be 2010; Najman et al., 2008; Yin
∼17.2 km from kinematic analysis of balance et al., 2010); Naga metamor-
sections. Thus, these fault-related folds have phic rocks (Aitchison et al.,
absorbed a minimum of 36.9 km of subhorizon- 2019; Najman et al., 2020); Tri-
tal shortening across the hinterland wedge (Text assic Pane Chaung turbidites
S1). In the study area, subhorizontal shortening (Sevastjanova et al., 2016; Yao
was perpendicular to the northwesterly trend et al., 2017); Burma Terrane
of the fold-and-thrust belts. This regime was and northern Mogok metamor-
also recorded by sets of conjugate NW-striking phic rocks (Zhang et al., 2018);
sinistral and NE-striking dextral faults, which Sibumasu Terrane (Burrett
­indicate principal stress distributions of subho- et al., 2014; Cai et al., 2017);
rizontal σ1 and σ3, and subvertical σ2 (Figs. 3 Tengchong Terrane (Li et al.,
and 4). More importantly, this contraction has 2016); Western Myanmar Arc
regional significance, as it has resulted in wide- (Licht et al., 2019; Zhang et al.,
spread folding and thrusting of Mesozoic to 2019b, and references therein);
Cenozoic strata, as well as the formation of a and West Sumatra (Zhang
young, 600-km-wide basin-orogenic belt in et al., 2019a).
western Myanmar (Brunnschweiler, 1966; Mau-
rin and Rangin, 2009a; Mitchell, 1993; Morley
et al., 2020; Zhang et al., 2018). The upper Mio-
cene to Pliocene Irrawaddy Group and equiva-
lents were the youngest strata involved in fold-
ing and thrusting (Kyi Khin et al., 2020; Pivnik
et al., 1998; this study), which implies that the
onset of the contraction may have occurred in the
late Miocene. Lithostratigraphic correlation on
flanks of the anticlines shown in P1, P2, P3, and
P6 indicates that at least 1.2–2.0 km of materials
have been eroded in the hinterland wedge (Text fold-thrust belt that reflects ongoing accretion parts of the southern Indo-Burman Ranges, the
S1 and Fig. S1). of fluvio-deltaic sediments above a décolle- outer wedge was more complex and exhibited
ment and varies in depth from 5 km to 10 km different geometries. Here, we divided the
The Outer Wedge, Offshore Southern (Betka et al., 2018; Bürgi et al., 2021; Maurin outer wedge into two tectonostratigraphic
Indo-Burman Ranges and Rangin, 2009a; Sikder and Alam, 2003). elements from west to east: the deforma-
The outer wedge recognized in the central This belt extends into the northern part of the tion front, which displays a series of closely
Indo-Burman Ranges and imaged by 2-D southern Indo-Burman Ranges and consists of spaced fault-related folds, and an inner sector
seismic reflection sections is a thin-skinned gentle antiforms (Figs. 5A and 5B). In other of the gentle shelf, which contains extensional

Geological Society of America Bulletin, v. 135, no. 9/10 2361

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Peng Zhang et al.

faults, thrust-top basins, and shale diapirs portion (off-scraped sediments) from the sedi- and declined toward the rear of the shelf terrace
(Figs. 5C–5G). ments that were underthrust beyond the base of after the late Paleogene.
Deformation front. The deformation front the deformation front and attached to the under- Shale diapirs. In the shelf terrace, where
includes three seaward-vergent imbricated side of the accretionary wedge (underplated). the thrusts are inactive or lacking, shale dia-
thrusts (FT1–FT3) with a regional topographic Shelf terrace. The shelf terrace (landward pirs dominate. Shale diapirs are associated
gradient of 5.7–9.8° and is 14–30 km wide, portion of the wedge) is 65–70 km wide at water with poorly defined zones of transparent or
extending from the base of the shelf edge at depths of 0–200 m and exhibits a surface slope chaotic reflections, and they display irregular
a depth of 200 m to the abyssal plain domain of only 0.2–0.4°. The thrusts in the shelf-terrace geometries. In several cases, outward-dipping
at depths of 2.2–3.2 km (Figs. 5C–5F). These domain are developed only within the Eocene reflectors in proximity to the main conduit were
thrusts were developed “in sequence” with the succession and are commonly steeper landward observed and traced discontinuously as far as
younger thrust (FT3) forming in the footwall of due to the back-rotation of thrust faults previ- the décollement (Figs. 5C–5E and S3). This
the older thrusts (FT2 and FT1) and produced a ously developed by ongoing frontal accretion. suggests that these shale diapirs were squeezed
forward-breaking sequence in the outer wedge. We interpret the thrusts to have formed within out of Eocene overpressured sediments. On
At the wedge toe, where the abyssal plain section a décollement at ∼5.5–6.0 sTWT (6.6–7.1 km; Ramree and Cheduba islands, mud volcanoes
(>6.0 km) is initially deformed, offset sedimen- Figs. 5 and S3), although their locations are are widely distributed and have been interpreted
tary reflectors (Fig. 6A) indicate that FT3 was difficult to determine in several places due to to have originated from methane-rich deposits
likely generated by an updip rupture of the abys- imaging limitations of the seismic sections. The buried at depths of 5.6–8.8 km and paleotem-
sal plain sediments (Bengal Fan). These thrusts thrust-top basins between these thrusts and FT1 peratures of 117–184 °C (Moore et al., 2019).
have deformed the Eocene to lower Pliocene were asymmetric and had a sediment thickness Maurin and Rangin (2009a) suggested that
successions and appear to continue at depths of up to 6.2 km (Fig. 5G). Overlying the Eocene the progressive collapse of the volcanic cone
of 5.0–6.0 sTWT (6.2–7.1 km) into positions succession (U1 horizon), the Oligocene to Pleis- induced circular synclines on Ramree Island
of high-amplitude reflections, which are inter- tocene sequences were offset by NW-striking, (Fig. 3). Shale diapirs and mud volcanoes
preted to represent a detachment surface dipping SW-dipping listric-style growth faults, and hang- appear to have developed mostly near the Ram-
to the east at angles between 1.4° and 2.1°. In the ing-wall units within these sequences thickened ree and Cheduba islands; they cannot be iden-
southernmost part of the southern Indo-Burman toward the growth faults (see below). The strata tified from poor-resolution seismic sections in
Ranges (near the Ayeyarwady delta), the cross- between the U4 and U6 horizons were also cut by the central and southern parts of the southern
sections (Figs. 5G and 5H) show that the defor- a younger set of planar extensional faults that dip Indo-Burman Ranges.
mation front appears as a positive flower-like toward both the NE and SW directions (Fig. 6B). Extensional faults in the RBA4 area. A
structure, which is a typical marker of a strike- The planar faults were particularly well-formed detailed structural analysis of the 2-D seismic
slip–dominated regime. The upper Pliocene to in the southernmost part of the southern Indo- data allowed us to recognize three sets of exten-
Quaternary successions on the right-hand side Burman Ranges (Figs. 5G and 5H), where the sional faults within the upper 3 s of the data in the
of the deformation front (shelf domain) thicken planar fault system often offsets the top upper central part of the RBA4 area (Figs. 3, 6, 7, and
toward the first thrust (FT1), which has activity Miocene unconformity (U3 horizon). 8). These include NW-striking late Paleogene to
dating to the late Pliocene to the present day. Reflectors in the Oligocene to Quaternary early Pliocene faults (Fault Set 1), NE-striking
An intervening synclinal (thrust-top) basin successions that onlap the underlying sediments Neogene to present-day faults (Fault Set 2), and
(5–15 km in width) on the back limb of the sec- are specifically described in seismic profile WNW-striking Pliocene intra-formational faults
ond thrust (FT2) was imaged in several seismic 4171 (Fig. 6B), which displays sediment growth (Fault Set 3).
sections (Figs. 5E and 5F). This thrust-top basin packages overlying FT1. At least two truncation NW-striking, SW-dipping extensional faults
is infilled with Oligocene to Quaternary growth surfaces were identified in this thrust-top basin. (Fault Set 1) cut the Eocene to lower Plio-
successions (1.9–2.3 km in thickness) that thin In the regions immediately overlying the ero- cene successions and characteristically exhibit
to the top of the FT2-related anticline, indicat- sion surface are two packages of low-amplitude, hanging-wall strata thickened toward the faults.
ing a prolonged active thrust from the late Paleo- high-frequency, laterally continuous reflectors These late Paleogene–Pliocene faults range
gene to Quaternary. Below this thrust-top basin that exhibit a distinct progradational pattern from ∼5–25 km in strike length with offsets of
sequence are poorly imaged Eocene sediments, dipping to the west (successions of the lower up to 0.4 s (∼370 m) in the RBA4 area (Figs. 6
interpreted here as the proto-wedge sequence. Pliocene and Quaternary; Figs. 6B and 6C). and 8). A specific interpretation of this set of
The Quaternary succession displays closely Such progradational reflection architecture has faults was conducted from seismic profile 4187
spaced, seaward-dipping reflectors and blan- also been recorded in the Oligocene and Mio- (Fig. 6C). These faults have vertical offsets of
kets of the underlying deformed successions, cene successions. The progradational reflectors ∼300–400 m and offset the U2 to U4 horizons
which are cut by a set of small-scale normal within the Oligocene–Quaternary successions to produce a series of landward-dipping rotated
faults (Figs. 5C and 5E), or slips that dip toward onlap the back limb of the anticline and are blocks. These listric-style faults are interpreted
the trench over a shallow detachment surface directly associated with FT1 (Fig. 6B). Over- to have been rooted at ∼6.0 sTWT on a décol-
(Fig. 5G). To the southwest, sub-parallel and lat- all, the fanning growth packages of sediments lement level, which may represent the landward
erally continuous reflectors that appear parallel deposited in the thrust-top basin likely indicate extension of the main décollement from the
to the deep oceanic crust/attenuated continental periodic or continuous synsedimentary activ- deformation front. Onlap reflectors and trunca-
crust (Rangin and Sibuet, 2017) dominate the ity of FT1, which progressively tilted the basin tion terminations were identified at the U4 to U6
sediments of the abyssal plain region (Oligo- landward. Growth successions are hardly visible levels, which indicates that multiphase erosional
cene to Quaternary successions; Figs. 5C, 5E, on the right-hand side of the shelf terrace (Fig. events occurred in the shelf terrace region. The
and Fig. S3). A strong reflector at 4.8–5.6 sTWT S3), which likely implies nearly dormant thrusts. NW-striking faults were later segmented by a
(6.0–6.7 km; Fig. S3) is interpreted to be the Our findings indicate that strain is currently dis- series of NE-striking Neogene to present-day
detachment surface that separates the accreted tributed primarily across the deformation front faults (Fault Set 2).

2362 Geological Society of America Bulletin, v. 135, no. 9/10

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Accretionary wedge development at oblique convergent margins

The Neogene to present-day faults (Fault Set (Fig. 9A). Detrital zircons with U-Pb ages older M11, and M07) and two forearc samples
2) cut through the Eocene to Pliocene strata than 300 Ma are less common (6–30%); how- (M220 and M221) yield pooled AFT ages that
and terminate upward within the Quaternary ever, in several cases (M211, M25, MY-57A, are fairly close to their stratigraphic ages within
sequence (Fig. 7). These faults are planar and M23, and M11), they formed subordinate peaks the 2σ level of error. The degree of annealing
dip steeply up to 85° both to the SE and NW. at 450–650 Ma and 0.95–1.2 Ga. All of the of apatite grains in these samples cannot be
Several of these faults clearly exhibit flower-like samples yielded MDAs between 49.6 ± 3.4 Ma definitively determined due to the uncertainties
structures (Fig. 7). The Neogene to present-day and 41.2 ± 1.0 Ma (Table 1). The detrital zir- regarding the depositional ages of the south-
faults have maximum offsets of ∼900 m and are con U-Pb age spectra from the Eocene strata ern Indo-Burman Ranges rocks. The Miocene
aligned in four NE-SW corridors in the RBA4 formed in the hinterland wedge of the southern samples from the outer wedge (M08 and M12)
survey (FS1–FS4; Fig. 8). Within the Miocene Indo-Burman Ranges exhibit age distributions show pooled AFT ages that are older than the
to Pliocene successions, hanging-wall units that are remarkably comparable to those of the stratigraphic ages, which indicates that they
thicken toward these faults and constrain their Central Myanmar Basin (Chindwin, Minbu, and are unreset. Thermal history modeling exhibits
activity from (at least) the early Miocene. In Shwebo-Minwun subbasins; Figs. 9 and 10). that all of the hinterland wedge samples (M07,
addition to these dominant faults, a subgroup of Although the newly presented Oligocene–Mio- M02, M23, M25, M28, and M211) and the
smaller NE-striking faults was also developed cene samples (M07, M12, and M08) and half southern Minbu subbasin samples (M220 and
within the Pliocene succession, primarily in of the published samples (10TTN13, 10TTN20, M221) may have experienced peak tempera-
the footwalls of the larger Neogene to present- R16-3-4, and 10TTN09; Naing et al., 2014; tures of >125 °C (Table 1; Fig. 12). This implies
day faults. Najman et al., 2020) show predominant peak that apatite grains in these samples were totally
The WNW-striking faults (Fault Set 3) are pla- ages at ca. 50 Ma and 450–650 Ma and sub- reset during post-depositional burial and reheat-
nar and dip ∼75° to the NE and SW (Figs. 5, 6B, ordinate peaks at ca. 65 Ma, 90–105 Ma, and ing. If samples M02 and M07 were excluded
and 8). Most of these faults were formed in the 0.95–1.3 Ga, the remaining published samples because of the very small numbers of grains
Pliocene succession. A small number of faults show age spectra and relative abundances similar (≤10; Text S3), the remaining four hinterland
extend downward, offsetting the U4 horizon and to those of the Eocene samples (Fig. 9A). All wedge samples would indicate an early Mio-
terminating in the upper Miocene strata; several Oligocene and Miocene samples plot close to the cene uplift event that began at ca. 22–12 Ma.
of these faults extend upward to the Quaternary Eocene samples on the multidimensional scaling In the southern Minbu subbasin, thermal mod-
strata (Figs. 5 and 6B). These faults have an aver- (MDS) graph (Fig. 9C). eling indicates two exhumation events that
age trace length of ∼5.5 km and are commonly initiated at ca. 30 Ma and 12 Ma, respectively.
mapped between the FS2 and FS3 NE-striking Detrital AFT results The oldest median AFT ages of samples M08,
faults (Figs. 8B and 8C). M12, and M11 are greater than their strati-
Out of seven analyzed samples in the hinter- graphic ages (Table 1), which indicates that
Detrital Zircon U-Pb Ages land wedge, four (M23, M25, M211, and M28) these samples were only partially reset. Ther-
yield pooled AFT ages of between 7.65 + 1.8/– mal modeling shows that the Miocene samples
The detrital zircon age spectra from the 1.5 Ma and 21.6 + 4.9/–4.0 Ma (see Table 1). did not experience peak temperatures of >75
Eocene strata in the hinterland wedge (M11, Our detrital zircon U-Pb ages indicate that °C (Fig. 12). This is consistent with syndeposi-
M23, M25, M28, and M211) are internally depositional ages of these four samples cannot tional structures observed in the field (coral and
consistent and show features similar to those be younger than 43 Ma (Fig. 9A). Since the burrow; Fig. S1) that suggest the Miocene rocks
of the published data (Figs. 9A and 9C). All AFT ages are significantly younger than the were deposited in a shallow water environment
of the Eocene samples are characterized by a stratigraphic ages, it is highly likely that apatite before being rapidly raised at ca. 8–3 Ma.
prominent age signal of 90–105 Ma. In addi- grains in these samples were completely reset, The late Miocene–Pliocene uplift history of
tion, most of the samples have two subordinate which would indicate the initiation of exhuma- the outer wedge was also recorded by several
peak ages at ca. 65 Ma and 50 Ma; these peaks tion of the hinterland wedge. The remaining angular unconformities (U4, U5, and U6) on
become prominent in samples M11 and M211 three samples in the hinterland wedge (M02, our seismic sections (Figs. 6 and 7).

TABLE 1. DETRITAL APATITE FISSION-TRACK (AFT) DATA FROM THE SOUTHERN INDO-BURMAN RANGES AND SOUTHERN MINBU SUBBASIN
Sample Depositional Pooled AFT age Pooled Ns Mean length Oldest AFT Peak age Peak T EasyRo† BasinRo§ EasyDL#
no. age (Ma) (track) (μm), 1σ age (Ma)* (Ma) (°C) Median Median Median
(Ma) CI +95%/–95% median (2σ) median (2σ) median (2σ) (2σ) (2σ) (2σ)
M23 44.8–44.4 7.65 +1.8/–1.5 92 13.56 ± 0.21 9.00 (5.44) ≥14.50 (9.95) ≥167.79 (26.33) ≥1.32 (0.36) ≥1.04 (0.33) ≥1.33 (0.45)
M25 56.4–42.8 11.3 +3.1/–2.4 68 13.54 ± 0.21 13.00 (7.33) ≥18.00 (11.47) ≥176.82 (20.63) ≥1.46 (0.30) ≥1.23 (0.32) ≥1.56 (0.45)
M211 46.0–40.8 16.8 +4.5/–4.6 72 13.58 ± 0.38 19.00 (12.45) ≥22.00 (13.97) ≥131.19 (19.76) ≥0.79 (0.13) ≥0.65 (0.08) ≥0.65 (0.11)
M220 46.5–43.3 29.1 +3.5/–3.1 345 13.37 ± 0.12 28.00 (6.49) ≥34.00 (7.54) ≥164.88 (20.19) ≥1.25 (0.30) ≥0.92 (0.25) ≥1.26 (0.41)
M221 24.6–22.2 13.4 +3.4/–2.7 79 12.82 ± 0.22 12.00 (5.08) ≥13.50 (5.24) ≥125.28 (23.08) ≥0.73 (0.15) ≥0.62 (0.09) ≥0.59 (0.11)
M02 46.6–42.1 23.2 +10.3/–7.1 31 12.80 ± 0.33 23.00 (13.85) ≥23.00 (13.95) ≥163.42 (12.08) ≥1.22 (0.18) ≥0.89 (0.16) ≥1.20 (0.28)
M07 32.6–29.0 17.9 +9.2/–6.1 24 11.56 ± 0.44 15.00 (24.48) ≥14.00 (10.67) ≥170.76 (28.24) ≥1.38 (0.38) ≥1.12 (0.36) ≥1.38 (0.49)
M08 11.6–5.3 26.8 +5.1/–4.3 156 13.59 ± 0.27 52.00 (19.84) 3.00 (1.87) 71.03 (23.55) 0.40 (0.10) 0.41 (0.09) 0.37 (0.09)
M11 43.2–39.2 31.0 +7.8/–6.3 88 12.25 ± 0.19 95.00 (26.01) 23.00 (16.87) 108.75 (15.42) 0.66 (0.12) 0.57 (0.09) 0.54 (0.14)
M12 23.0–16.0 32.0 +6.8/–5.6 125 13.10 ± 0.12 95.50 (42.41) 9.00 (8.19) 73.23 (19.01) 0.42 (0.08) 0.43 (0.07) 0.39 (0.07)
M28 46.0–40.0 21.6 +4.9/–4.0 108 13.07 ± 0.32 18.00 (13.30) ≥19.00 (14.14) ≥177.03 (14.90) ≥1.45 (0.24) ≥1.23 (0.29) ≥1.56 (0.40)

*If the Oldest AFT Ma median (2σ) is less than the stratigraphic age, the sample has been reset.
†EasyRo from Burnham and Sweeney (1989).
§BasinRo from Nielsen et al. (2017).
#EasyDL from Schenk et al. (2017).

Geological Society of America Bulletin, v. 135, no. 9/10 2363

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Peng Zhang et al.

Figure 12. Plots show thermal history modeling for samples collected from the southern Indo-Burman Ranges and southern Minbu sub-
basin, using program FT3. Note that all temperature-time (t-T) paths are shown, and all—except for the first t-T path that fits the apatite
fission-track (AFT) data—are only shown back to the age of the oldest fission track present for each subgroup. Full results and specific
analyses are shown in Text S3 (see footnote 1).

DISCUSSION the deformation front (FT1; Fig. 6A), which analysis of the progradational reflectors within
caused a narrow, steep deformation front (aver- the Miocene to Quaternary successions that
Kinematic Evolution of the Outer Wedge age width of 15.6 km) and broad, low-relief straddle the depocenter of the thrust-top basin
Under Highly Oblique Plate Convergence shelf terrace (average width 49 km). Because and terminate toward FT1 (Fig. 6B). The vertical
of the steeper surface slope of the deformation stack of progradational reflectors strongly indi-
Our data are the first direct records of wedge front (average 7.4°) and lower width ratio of the cates that the sedimentation rate may have never
deformation in the southern Indo-Burman deformation front within the entire outer wedge exceeded the activity rate of the growing thrust
Ranges, and they yield several significant char- (average 0.24), the geometry of the outer wedge during most geological stages. Periodic slip on
acteristics that may provide insights into global seems to be unique to this margin when com- this thrust fault thus produced the landward-
commonalities and site-specific characteristics pared to other convergent margins such as Suma- tilted slope-basin successions at the rear of the
of the tectonic evolution of accretionary margins. tra (Indonesia), Nankai (Japan), and Hikurangi deformation front. The remaining thrust faults
A first-order characteristic feature is the con- (New Zealand; Table 2). The effect of this long- (FT2 and FT3) in the deformation front are prob-
tinuous activity of the thrust fault at the rear of lived thrust fault is evident through qualitative ably inactive or were initiated very late, because

2364 Geological Society of America Bulletin, v. 135, no. 9/10

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Accretionary wedge development at oblique convergent margins

TABLE 2. COMPARISON OF STRUCTURAL PARAMETERS CALCULATED FOR THE SOUTHERN INDO-BURMAN RANGES
(RAKHINE) FROM THIS STUDY, WITH PUBLISHED DATA FROM SELECTED OTHER MARGINS
Represented Max trench Surface slope of Surface Wedge taper values (°) Deformation Shelf Ratio Obliquity
margins* sediment the deformation slope of the front width terrace L1/(L1+L2) (°)†
thickness front shelf terrace Surface Basement Taper (L1; km) width
(km) (α1; °) (α2; °) slope slope (β) (α+β) (L2; km)
(α)
Rakhine, Myanmar >6.0 7.4 0.2 1.9 2.1 4.0 15.6 49.0 0.24 76
Northern Sumatra, Indonesia 4.0 3.3 0.4 1.3 1.6 2.9 53.3 43.0 0.57 53
Southern Sumatra, Indonesia 1.5 4.7 0.3 2.1 2.2 4.3 30.0 64.0 0.32 31
Eastern Java, Indonesia 1.4 5.8 0.9 3.2 6.7 9.9 38.1 44.1 0.46 5
Nankai, Japan 1.6 5.1 0.1 1.9 4.6 6.5 29.0 50.0 0.37 30
Honshu, Japan 0.3 5.4 0.8 2.9 5.0 7.9 48.0 55.0 0.47 33
Aleutian, USA 1.0 2.6 0.1 1.0 4.2 5.2 15.4 32.5 0.32 20
Alaska, USA 2.2 3.3 0.2 1.8 2.4 4.2 80.0 76.0 0.51 25
Cascadia, USA 2.9 2.6 0.3 2.0 1.8 3.8 87.0 68.0 0.56 44
Colombian-Caribbean 4.0 4.3 0.2 2.2 2.1 4.3 78.0 71.0 0.52 19
Lesser Antilles 5.0 3.5 0.4 2.0 1.0 3.0 65.0 215.0 0.23 28
Central Chile 2.0 4.0 0.3 2.4 7.6 10.0 52.0 40.5 0.56 26
Southern Chile 3.0 4.9 0.7 2.8 3.4 6.2 44.7 42.0 0.52 48
Makran, Pakistan 6.0 2.8 1.3 2.0 1.5 3.5 72.5 73.6 0.50 9
Southern Hikurangi, New Zealand 3.8 3.2 1.4 1.0 3.0 4.0 93.8 31.3 0.75 50–80
*Key References: Rakhine (this study); Sumatra (McNeill and Henstock, 2014; Moeremans et al., 2014); Java (Lüschen et al., 2011); Nankai (Moore et al., 2007, 2015;
Strasser et al., 2009); Honshu (von Huene et al., 1994); Aleutian (von Huene and Scholl, 1991); Alaska (von Huene and Klaeschen, 1999); Cascadia (Adam et al., 2004);
Colombian-Caribbean (Mantilla-Pimiento et al., 2009); Lesser Antilles (Bangs et al., 1990); Central Chile (Contreras-Reyes et al., 2010); Southern Chile (Polonia et al.,
2007); Makran (Grando and McClay, 2007; Smith et al., 2012); Hikurangi (Barnes and de Lépinay, 1997; Barnes et al., 2010).
†Data are summarized from Noda (2016), Barnes and Lépinay (1997), and Mallick et al. (2020).

only ∼5 km of horizontal shortening (measured steeper faults with larger displacements occur have controlled the Indo-Burma Subduction
on the U1 horizon) has occurred on these faults in the transpressional wedge, which absorbs the Zone since at least the Miocene (Bertrand and
over the past ∼30 m.y. The thrust faults that majority of the horizontal shortening. The slip Rangin, 2003; Betka et al., 2018; Bürgi et al.,
occur between 17°00′N and 19°30′N show steep on these thrusts is comparatively long-lived, 2021; Mallick et al., 2020; Maurin and Rangin,
dips at their bases and are less steep toward their and thus the development of the next thrust fault 2009a; Morley et al., 2020; Pivnik et al., 1998;
tops with a convex-up geometry (Fig. 5C–5F), is delayed. Rangin et al., 2013; Vigny et al., 2003). Global
while those faults that occur between 15°50′N A second observation was the occurrence positioning system measurements predict a
and 17°00′N are arranged as positive flower of NE-striking extensional faults (Fault Set 2). relative India–Sundaland motion of 36 mm/yr
structures (Figs. 5G and 5H). These planar faults with steeper dip angles often oriented N014 at latitude 17° N (Socquet et al.,
Although the geometric and structural prop- cut through entire wedge successions and pro- 2006). Considering the 18–23 mm/yr north–
erties of the deformation front in the southern duce a series of negative flower-like structures south motion of the West Burma Terrane with
Indo-Burman Ranges differ from those observed (Fig. 7). In the RBA4 region, these extensional respect to Sundaland on the Sagaing Fault in the
at other convergent margins, these features are faults are perpendicular to the frontal thrust east, a rest 7–9 mm/yr of this motion would be
largely compared to analog modeling in the lab- faults and the NW-striking listric-style faults at expected to be distributed in the Indo-Burman
oratory under highly oblique plate convergence the rear of the shelf terrace (Fault Set 1), split- Ranges (Mallick et al., 2020; Maurin et al.,
(Leever et al., 2011; McClay et al., 2004). These ting the study region into five blocks with dif- 2010; Vigny et al., 2003).
theoretical studies show that a larger obliquity, ferent directions of relative motion (Fig. 8D). The presence of shale diapirs and mud vol-
for example φ >60°, would produce a rather For example, block A moved northeast along canoes is another characteristic associated with
symmetrical flower-like structure with steep FS1 relative to block B, whereas block C the development of an accretionary wedge in the
faults rooting near the basal décollement and moved southwest along FS3 relative to block southern Indo-Burman Ranges. These fluid-sed-
a decreased number of large-offset prowedge D. These characteristic features are comparable iment mixtures, along with a lower wedge taper
thrust faults. In this context, the accretionary to those of Riedel shears in a right-lateral shear (4.0°; Table 2) and a thick incoming sediment
wedge experienced dramatic vertical uplift in system (Fig. 8D). Because the distinct wedge- section (>6 km; Bengal Fan), indicate a fluid-rich
the axial belt and limited forward expansion. shaped reflectors first occurred in the Miocene margin of the southern Indo-Burman Ranges.
The geometrical characteristics of the transpres- sequences, we propose that the right-lateral This margin would have a weak basal surface,
sional wedge vary with the convergence obliq- shear motion in the Indo-Burma Subduction which commonly presents as a high-amplitude,
uity due to the changing orientation of the stress Zone initiated in the Miocene and continues to negative-polarity reflector on seismic profiles
triad (Burbidge and Braun, 1998; Koons, 1994; the present day. This is roughly backed by our that reflects reduced bulk density, increased
Leever et al., 2011). We define a reference frame field measurements, which demonstrate that porosity, and probably high pore fluid pressures
in which x and y are both horizontal and per- the southern Indo-Burman Ranges have been (e.g., Bangs et al., 2004). However, the décol-
pendicular and parallel to the plate boundary, reshaped by dextral strike-slip motion since the lement level recorded in our seismic reflection
respectively, and z is vertical. For orthogonal Miocene (Fig. 3). data was unreflective and/or discrete (Fig. S3).
convergence, σ1//x, σ2//y, and σ3//z, whereas in Along the Sunda trench between 14°00′N and This finding could be explained by either its spe-
strike slip, σ1 and σ3 are horizontal at an angle 16°30′N (Fig. 1), Nielsen et al. (2004) reported cific properties, which do not generate distinct
of 45° to x and y, and σ2//z (Burbidge and Braun, high-quality, 2-D seismic and swath bathymet- reflectors on the seismic profiles, or a sub-critical
1998). Thus, σ1 rotates horizontally around the ric data, in which positive flower structures were taper model, in which the observed overpressure
x-axis with increasing convergence obliquity, recognized in seismic cross sections, and dex- is dispersed over the accreted sediment section
which in turn changes the geometry of the thrust tral strike-slip faults were mapped on the plan rather than concentrated at the décollement
faults in the deformation front. For instance, graph. In fact, dextral strike-slip tectonics may surface (Smith et al., 2012). The occurrence of

Geological Society of America Bulletin, v. 135, no. 9/10 2365

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Peng Zhang et al.

shale diapirs and mud volcanoes in the offshore the southern Indo-Burman Ranges and constrain ers that flowed through the eastern Himalayas
and onshore Oligocene to Quaternary succes- sedimentation spanning the early to middle and SE Tibetan Plateau margin (northern Mogok
sions (Figs. 3 and 5) suggests that the observed Eocene (50–41 Ma). Paleontological data from Metamorphic Belt and Dianxi-Burma batho-
overpressures were diffused upwards. Therefore, the hinterland wedge (Lei et al., 2009; Naing, liths; Fig. 1), respectively. The formation of the
the outer wedge of the southern Indo-Burman 2019) are broadly consistent with the MDAs, ­palaeodrainage coincided with initial uplift of
Ranges seems to have developed as a subcritical which implies that the youngest detrital zircon SE Tibetan Plateau in the late Eocene–early Oli-
wedge with a stronger décollement than origi- U-Pb age populations can be used to constrain gocene (Hoke et al., 2014; Su et al., 2019; Xiong
nally thought; the relatively strong décollement the maximum stratigraphic ages. Naing (2019) et al., 2020; Xu et al., 2015). A late Eocene–early
could sustain vertical growth of the outer wedge. reported Late Cretaceous planktonic foraminif- Oligocene unconformity (Licht et al., 2019; Naj-
However, this effect appears to be limited due eral assemblages and nannofossils and Paleo- man et al., 2022; Westerweel et al., 2020; Zhang
to the lack of significant plate subduction in cene to early Eocene radiolarian fauna (Fig. et al., 2017) discovered in the Chindwin subba-
this region, as inferred from previous seismic S2) from a few isolated observation sites along sin marks a significant erosional event that may
tomography (Pesicek et al., 2010; Replumaz the Pyay–Toungup section (hinterland wedge). have caused sediment reworking of the Yaw For-
et al., 2010). Our seismic reflection data (Fig. However, differentiating the Upper Cretaceous mation and subsequent deposition in the south-
S3) show that only ∼1-km-thick Eocene rocks to Paleocene from the Eocene strata is difficult ern Indo-Burman Ranges.
were carried down the trench before the late because of the heavily forested nature of the Our analysis of the Miocene rocks exposed
Pliocene, and they triggered tectonic underplat- hinterland wedge. Consequently, knowledge of on Ramree Island and along the Rakhine coast
ing and subsequent NE–SW-striking extensional the Upper Cretaceous to Paleocene rocks is still (Fig. 3) refutes the earlier hypothesis of a domi-
faulting (Fault Set 1). limited, although the Eocene Rakhine Flysch is nant Himalayan source (Allen et al., 2008;
Our analysis reveals that the outer wedge more prevalent in the study region. ­Curray, 2005; Naing et al., 2014). The typical
of the southern Indo-Burman Ranges has The Eocene samples of the southern Indo- Himalayan-derived detritus preserved in the
­conspicuous geometric and structural character- Burman Ranges display detrital zircon age spec- Miocene foreland depression (Sub-Himalaya
istics that potentially correlate with a transpres- tra similar to those from the Burmese forearc and Bengal Basin) and submarine fans (Bengal
sional wedge that formed under highly oblique subbasins (i.e., Chindwin and Minbu) and cen- and Nicobar) is characterized by Proterozoic to
convergence. However, there are also marked tral Indo-Burman Ranges but substantially dif- early Paleozoic zircons (Fig. 10A; Blum et al.,
differences between the outer wedge of the ferent age spectra from those of the northern 2018; Bracciali et al., 2015; McNeill et al., 2017;
southern Indo-Burman Ranges and the theo- Indo-Burman Ranges. The strata of the north- Najman et al., 2008, 2012; Pickering et al., 2020;
retical results, as well as several natural wedges ern Indo-Burman Ranges consist primarily of Webb et al., 2021). However, the Miocene rocks
at highly oblique convergent margins such as Proterozoic to early Paleozoic age populations in the southern Indo-Burman Ranges are domi-
Sumatra (φ = 59–37°; McNeill and Henstock, with major peaks at 470–600 Ma and 0.8–1.2 Ga nated by Cretaceous–Cenozoic detrital zircons
2014; Moeremans et al., 2014), Hikurangi and subordinate peaks at 1.5–2.0 Ga (Fig. 10B; (Fig. 9A), while Himalayan age signals (peaks
(φ = 40–10°; Barnes and de Lépinay, 1997; Aitchison et al., 2019; Ding et al., 2022; Vadlam- at 450–650 Ma, 0.8–1.2 Ga, and 1.60–2.0 Ga;
Barnes et al., 2010), and Cascadia (φ = 46°; ani et al., 2015), which are typical detrital zircon Fig. 11) are either minor or lacking (11.6–28.7%).
Adam et al., 2004), which have different defor- age signals of the Himalayan Orogen (Fig. 11). The MDS in Figure 10E compares data sets from
mation front widths, fault types, and thicknesses Although the contributions from the Gangdese the Indo-Burman Ranges, Central Myanmar
of slope-basin infillings. This is hardly surpris- Arc (Robinson et al., 2014) and/or the Himala- Basin, Himalayan foreland basins, submarine
ing, given that the formation of an accretion- yan Orogen (Arboit et al., 2021) to the Burmese fans, and modern river sand. The Miocene rocks
ary wedge is often affected by various factors, forearc subbasins in a certain period time of the clearly plot within the West Burma Terrane field
including the material properties of the wedge Eocene are not completely eliminated, previous and show a closer affinity to the lower–middle
and the décollement (friction, cohesion, and studies (Cai et al., 2020; Licht et al., 2013, 2019; Eocene rocks deposited in the Burmese forearc
pore fluid pressure), plate convergence (obliq- Wang et al., 2014; Zhang et al., 2019b, 2021a) subbasins and the southern Indo-Burman Ranges
uity and velocity), isostatic response (uplift and have conservatively linked the Upper Cretaceous itself. Given that (1) there are no obvious uncon-
subsidence), and surface processes (erosion and to middle Eocene rocks in these subbasins to the formities between the lower to middle Eocene
sedimentation; Noda et al., 2020; Tobin and Saf- exhumation and erosion of autochthonous vol- rocks and their overlying strata in the deeper
fer, 2009; von Huene and Ranero, 2003). Since canogenic rocks, granitoids, and deep basement parts of the Burmese forearc subbasins (Pivnik
at least the Miocene, increased trench sedi- rocks of the West Burma Terrane (wedge core et al., 1998; Zhang et al., 2017), and (2) erosion
mentation rate and enhanced hinterland wedge of the Indo-Burman Ranges, Wuntho-Popa Arc, of the western outcrops in the Miocene would
­erosion have been observed in the southern Indo-­ and Tagaung-Myitkyina Belt; Fig. 1). Our find- result in mixed age signals that correspond to
Burman Ranges (McNeill et al., 2017; Najman ings, when combined with previously published both the Burmese forearc subbasins and wedge
et al., 2012, 2020; Pickering et al., 2020; this detrital zircon data, strongly suggest that the core of the central Indo-Burman Ranges (Trias-
study); however, the influence of these processes southern Indo-Burman Ranges was open to the sic Naga metamorphic rocks and Pane Chaung
on wedge growth remains unknown. We address trench to the west (in modern coordinates) and turbidites; Figs. 1, 11, and S4; see footnote 1),
this question below. represented an additional significant discharge we propose that the Miocene rocks are a result of
region for material derived from West Burma sediment reworking of the hinterland wedge of
Reworking of the Hinterland Wedge Rocks Terrane during the early–middle Eocene. The the southern Indo-Burman Ranges. On another
Facilitates Vertical Growth of the Outer Oligocene rocks show detrital zircon age pat- MDS graph (Fig. 9C) that shows only the south-
Wedge terns very similar to those of the upper Eocene ern Indo-Burman Ranges and southern Minbu
rocks in the Chindwin subbasin (Yaw Forma- data sets, the Miocene rocks plot close to the
Detrital zircon U-Pb ages provide MDAs for tion; Fig. 10E), which reflects the influence of Eocene samples and significantly overlap, which
sediments exposed in the hinterland wedge of both the proto-Siang and Upper Irrawaddy riv- indicates that they have strong affinity.

2366 Geological Society of America Bulletin, v. 135, no. 9/10

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Accretionary wedge development at oblique convergent margins

The outer-wedge rocks derived from the hin- at a shift from southwestward transverse flows the Miocene (Khin et al., 2020; Morley et al.,
terland wedge require that the latter has been into the Bay of Bengal in the Paleogene (Licht 2020; Najman et al., 2020; this study). Detritus
uplifted since the early Miocene. This is sup- et al., 2013, 2019; Thein and Maung, 2017) to derived from the Minbu subbasin might have
ported by syndepositional structures observed a southward longitudinal flow into the Prome- been transported to the Andaman Sea by the
at the Rakhine coast (Fig. S1) and thermal his- Irrawaddy subbasin and Andaman Sea in the through-going Irrawaddy River (Jonell et al.,
tory modeling of our AFT data. Four samples Neogene (Jonell et al., 2022; Khin and Myitta, 2022; Zhang et al., 2021a). Our MDS graph
from the hinterland wedge (M23, M25, M28, 1999; Thein and Maung, 2017). Second, various (Fig. 10E) shows that the Miocene rocks plot
and M211) were buried to depths greater than Eocene nannofossils are found in the lower Mio- closer to the Eocene hinterland wedge than to
the partial annealing zone and showed fully reset cene sedimentary rocks (Najman et al., 2020). the contemporaneous rocks from the Minbu or
AFT ages, which suggests that initial uplift of Third, truncation reflectors that record Neogene Chindwin subbasin. We further propose that sed-
the hinterland wedge occurred at ca. 22–12 Ma erosional events at the rear of the outer wedge iments eroded from the hinterland wedge must
(Fig. 12). If we assume a standard geothermal were identified in our 2-D seismic sections have been transported along the footwalls of the
gradient of 30 °C/km and put the base of the (Figs. 5–7). NE-striking transextensional faults (Fault Set 2),
partial annealing zone at ∼120 °C, average Large-scale progradational reflectors where NE-striking elongated depocenters have
exhumation rates of the hinterland wedge can be (8–16 km in length) within the Oligocene to been observed (Fig. 8). Our findings suggest
estimated at 182–333 m/m.y.; at these relatively Quaternary sequences provide unambiguous that transextensional faults under oblique plate
steady exhumation rates, the Eocene rocks of the evidence of stable southwestward flow into convergence governed the pathways of reworked
hinterland wedge were raised to depths less than the shelf terrace of the outer wedge (thrust-top sediments from the hinterland to the outer wedge
those of the partial annealing zone (∼2 km) at ca. basins; Figs. 6B and 6C). One could argue that (Fig. 13).
10–5 Ma (Fig. 12 and Text S3). This implies that the presence of progradational reflectors within This study lacks robust detrital zircon age
the hinterland wedge of the southern Indo-Bur- the outer wedge may be directly linked to the constraints on sediments in the deformation
man Ranges must have been partially exposed to continuous sediment supply from the Central front because systematic sampling is currently
sub-aerial condition, if the overlying Oligocene Myanmar Basin (Minbu subbasin) to the east unavailable. The immature thrust-top (wedge
rocks with a thickness of 1–2 km were included by trans-wedge fluvial networks. This is most slope) basin lying between FT1 and FT2
(Maurin and Rangin, 2009a; Naing et al., 2014; likely true for Oligocene sedimentary rocks, as (Figs. 5E and 5F) shows a clear progradational
Zhang et al., 2021a). In addition, the uplift of the the Minbu subbasin was characterized by the reflection configuration in the Quaternary suc-
hinterland wedge in the Miocene is further sup- south–southwestward direction of rivers dur- cession that extends laterally into the thrust-top
ported by three independent lines of evidence. ing the Oligocene, and these rivers could thus basin in the shelf terrace. This indicates that the
First, an increase in sedimentation rate in the feed sediments to offshore basins to the south- reworked sediments from the hinterland wedge
Prome-Irrawaddy subbasin started in the late west (Gough et al., 2020; Zhang et al., 2021a). may have been transported in the deforma-
Miocene, with a large N-S–striking submerged However, substantial discharge from the Minbu tion front by submerged distributary channels
delta occurring in the Andaman Sea (e.g., Racey subbasin thereafter is unlikely because the hin- or large-scale slumping of surficial sediments
and Ridd, 2015; Zhang et al., 2021a). This hints terland wedge has emerged above sea level since near the shelf break (Figs. 5 and 13). However,

Figure 13. Schematic shows de-


velopment of the outer wedge
under highly oblique plate
convergence during Neogene
to Quaternary times. The de-
nuded region represents the up-
lifted hinterland wedge, where
trans-wedge fluvial networks
transport the reworked hinter-
land wedge materials into the
slope-basin along the NE–SW-
striking transtensional faults
(Fault Set 2) to form large sea-
ward progradational sequences
in the outer wedge. While most
of these materials were trapped
by the slope-basin due to con-
tinuous fault slip on the thrust
at the rear of the deformation
front (FT1), a small amount of
the eroded materials was deliv-
ered into the deformation front,
where they were deformed along
with the abyssal plain sediments
(Bengal Fan). The black arrow
represents the motion vector of the Indian plate and indicates a highly oblique convergent margin of the southern Indo-Burman Ranges.

Geological Society of America Bulletin, v. 135, no. 9/10 2367

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Peng Zhang et al.

the imbricated thrust faults that nucleated “in Ridge and West Burma Terrane. However, the sloping accretionary wedge fed by transverse
sequence” (Figs. 5C–5H) and the abyssal plain smooth reflectors of the abyssal plain succes- rivers draining through the West Burma Terrane
sediments offset by the youngest frontal thrust sions observed at the wedge toe (Figs. 5 and S3) (Licht et al., 2019; Najman et al., 2020; Zhang
(FT3; Fig. 6A) strongly imply that Bengal Fan indicate that the ridge may not have arrived at et al., 2019b). The effect of the collision between
sediments accreted into the deformation front the base of the wedge yet. Our findings high- India and West Burma Terrane propagated to
and were probably the dominant sources. light the significance of the highly oblique plate the central Indo-Burman Ranges in the middle
In summary, our analyses demonstrate that convergence and sediment reworking that trig- Eocene (45–40 Ma; Morley et al., 2020; Najman
the hinterland wedge rocks are dominantly gered the vertical expansion of the outer wedge et al., 2022), resulting in an initial uplift of the
derived from the West Burma Terrane, while during the Miocene to Quaternary. Based on its wedge core and forward growth of the hinterland
the shelf terrace rocks are the product of Eocene structural styles and provenance characteristics, wedge (Fig. 14A). Because of the smaller obliq-
hinterland wedge sediment reworking, and the the evolution of the entire southern Indo-Burman uity (φ< 50°; Mallick et al., 2020), the forward
deformation front rocks are sourced from the Ranges can be summarized in two major stages growth of the outer wedge in the central Indo-
Himalaya (Bengal Fan) with a subordinate (Fig. 14). Burman Ranges appears to be mainly controlled
contribution from the hinterland wedge. These Stage I: Deposition of turbidites (Rakhine by frontal accretion of fluvio-deltaic sediments
findings are of great importance for understand- Flysch) and forward expansion of the proto- in the trench rather than by oblique plate con-
ing the development of accretionary wedges at wedge. The study region in the Eocene was a vergence. It is not clear whether this process
other highly oblique plate convergent margins, seaward-sloping accretionary wedge in which is influenced by the sediment reworking of the
particularly the outer wedge. braided channel and/or submarine fan deposi- overriding plate material. We propose that even
According to the critical-taper theory (Dahlen tion occurred. These turbidites were deformed if this influence exists, it would be fairly minor
et al., 1984; Davis et al., 1983), surface erosion “in sequence” due to consistent frontal accretion, because sediments in the outer wedge were pre-
of the hinterland wedge and fast sedimenta- which caused subsequent underplating of the dominately supplied by the Himalayan Orogen
tion in the thrust-top basins of the outer wedge Eocene turbidites at the base of the proto-wedge and Gangdese Arc to the north (Fig. 10). Ongo-
decrease the topographic slope and produce a (Fig. 14A); in addition, underplating of sedi- ing frontal accretion during the past 8 m.y. pro-
subcritical taper, which requires vertical fron- ments thickened and uplifted the proto-wedge. duced a wide (maximum width of 200 km) and
tal deformation to maintain an overall critical Between the pro-vergent thrusts, the Oligocene thin-skinned fold-thrust belt (Figs. 14B and 14C;
taper. This partly explains the occurrence of a rocks derived from the recycling of the upper Betka et al., 2018; Bürgi et al., 2021; Maurin and
narrow deformation front in the study region, Eocene forearc basin rocks were deposited in Rangin, 2009a).
which had higher slope angles and limited fron- structural lows. The wedge development at this We further note that at least three additional
tal accretion (Figs. 5C–5H and 6A). Previous stage was not significantly affected by oblique mechanisms could explain the forward accretion
numerical and analog studies demonstrated that plate convergence. of the outer wedge in the central Indo-Burman
erosion of an accretionary wedge would pre- Stage II: Transpressional wedge development Ranges, which are described below.
vent the initiation of new thrusts in the defor- under highly oblique convergence and sediment (1) By altering the relative balance between
mation front and instead concentrate strain on reworking. Since the early Miocene, the study the rate of material input to the wedge and the
these existing active thrusts within this belt region has developed as a highly oblique con- forces resulting from gravity acting on the ele-
(Cruz et al., 2010, 2011; Mugnier et al., 1997). vergent margin as a result of plate reorganiza- vation contrast between the hinterland wedge
Behind these existing active thrusts, additional tion in the northeastern Indian Ocean (Bertrand and the outer wedge (Ball et al., 2019; Copley
basal shear stress related to active sedimentation and Rangin, 2003; Curray, 2005; Rangin et al., and McKenzie, 2007; Najman et al., 2020), an
strengthens the coupling of the interface between 2013). This led to substantial vertical uplift in the increase in the thickness of the incoming sedi-
the wedge base and décollement (Burbidge and hinterland wedge and an increase in fault activity ment section contributed to obvious changes in
Braun, 1998; Noda et al., 2020). This behavior on the thrusts at the rear of the deformation front the uplift and deformation of the outer wedge.
is similar to the highly oblique plate conver- (Fig. 14B). The eroded materials from the hinter- In front of the wedge, the subducting Cenozoic
gence described above. Therefore, it seems that land (>1.2–2.0 km in thickness) were primarily strata in the Bengal Basin have a maximum
higher convergence obliquity, surface erosion, trapped by slope basins in the shelf terrace, fur- thickness of >16 km (Curray, 1991), which
and active sedimentation increase the duration ther strengthening the vertical slip on the thrusts is larger than any other convergent margin in
and magnitude of slip along individual thrusts at the rear of the deformation front. In this stage, the world.
in the deformation front, particularly along FT1 frontal accretion and underplating of trench sedi- (2) The initial subduction of the negatively
(Figs. 5 and 6). Thus, periodic activity on these ments below the wedge were limited. buoyant Indian passive margin lithosphere into
thrusts produces a greater fault dip and leads to In contrast to the southern Indo-Burman the mantle produced increased flexural loading
more accommodation space, which allows the Ranges, the central Indo-Burman Ranges (20– in the Surma subbasin (northeast Bengal Basin;
development of growth sequences in thrust-top 25°N) have an additional Mesozoic wedge core, Fig. 1) and steepened the topographic slope of
basins (Leever et al., 2011). a more strongly deformed hinterland wedge, the outer wedge (Mallick et al., 2020). As the
and a broader folded and thrusted outer wedge wedge was perturbed from its critical taper by
Differential Evolution between the (Fig. 14C; Betka et al., 2018; Bürgi et al., 2021; sinking, it was driven to propagate westward to
Southern and Central Parts of the Maurin and Rangin, 2009a; Morley et al., 2020). reduce its taper and sustain equilibration. The
Indo-Burman Ranges These differences reflect distinct kinematic and subduction of the Indian lithosphere beneath the
dynamic evolutionary processes. Prior to the West Burma Terrane has been well documented
Maurin and Rangin (2009b) attribute the early Eocene (48 Ma), the tectonic and sedi- in seismic tomography images (Pesicek et al.,
oversteepened deformation front and fairly mentary settings in the central Indo-Burman 2010; Replumaz et al., 2010).
flat shelf terrace of the southern Indo-Burman Ranges were comparable to those in the southern (3) The Tibetan Plateau crustal flow rotated
Ranges to the collision between the Ninetyeast Indo-Burman Ranges; both areas had a seaward- southwestward around the Eastern Himalayan

2368 Geological Society of America Bulletin, v. 135, no. 9/10

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Accretionary wedge development at oblique convergent margins

Figure 14. Paleogeographic reconstructions of the West Burma Terrane (WBT) are shown with schematic of cross section
across W–E transects in the central and southern parts of the Indo-Burman Ranges (IBR). (A) The WBT was dragged north-
ward by the Indian plate and dominated by trans-wedge rivers until the middle Eocene uplift of the central and northern
parts of the Indo-Burman Ranges. Prior to this period, both the central and southern parts of the Indo-Burman Ranges had
a broad deformation front characterized by numerous folds and thrusts. (B) The WBT has collided with Asia since at least
the early Oligocene and became a highly oblique convergent margin in the early Miocene. Highly oblique plate convergence
and sediment reworking of the hinterland wedge resulted in continuous fault slip on thrusts at the rear of the deformation
front and delayed forward advance of the wedge. (C) Differential evolution of the southern and central parts of the Indo-
Burman Ranges controlled different wedge geometry and structure. The black arrow represents the motion vector of the
Indian plate. Tectonic reconstructions are modified after Westerweel et al. (2019) and Morley et al. (2021). Cross section of
the central Indo-Burman Ranges is from Maurin and Rangin (2009a). IR—Irrawaddy River.

Geological Society of America Bulletin, v. 135, no. 9/10 2369

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Peng Zhang et al.

Syntaxis was initiated in the middle Miocene uplift and slope steepening by concentrating zontal rotation of the maximum compressive
(Clark and Royden, 2000; Liu et al., 2014) and stress on fracture zones immediately above the stress (σ1), which changes the geometry of
caused uplift of the inner wedge (hinterland seamounts (Ruh, 2016; Wang and Bilek, 2011), the thrust faults in the deformation front, par-
wedge and Mesozoic wedge core) and collapse leading to large-offset, long-lived thrust faults ticularly the oldest thrust fault at the rear (FT1).
of the outer wedge (Rangin et al., 2013; Rangin and a delay in the generation of new thrusts in The slip on FT1 is comparatively long-lived,
and Sibuet, 2017). The southwestward Tibetan front of the seamounts (e.g., Morgan and Bangs, and thus the formation of the next thrust faults
Plateau crustal flow was restrained between the 2017). At nonaccretionary-type margins where (FT2 or FT3) is delayed. However, the features
Shillong Plateau and the Ninetyeast Ridge; thus, a narrow deformation front and limited trench of the outer wedge are not completely coincident
only the outer wedge in the central Indo-Bur- sediments occur, subduction erosion can desta- with theoretical results and those observed from
man Ranges shows a wider fold-and-thrust belt. bilize the upper plates and tip them seaward, other highly oblique convergent margins, such
However, this crustal flow arrived at the Indo- thereby forming oversteepened slopes (Noda, as Sumatra, Hikurangi, and Cascadia. Our detri-
Burman Ranges later than the timing of initial 2016; von Huene and Ranero, 2003). However, tal zircon U-Pb and apatite fission-track analy-
uplift of the central Indo-Burman Ranges, which the effect of highly oblique convergence or sedi- ses indicate that the majority of the sediments
was dated to the middle Eocene by recent studies ment reworking has been studied in detail at nei- blanketing the slope basins within the shelf ter-
(Morley et al., 2020; Najman et al., 2022). ther accretionary-type margins (e.g., Sumatra, race are the product of sediment reworking of
Longitudinal differentiation has been a pri- Nankai, and Hikurangi) nor nonaccretionary- the hinterland wedge since the early Miocene
mary characteristic of the Indo-Burma Sub- type margins (e.g., Eastern Java, Central Chile, (22–12 Ma). Active sedimentation behind the
duction Zone since at least the Miocene and and Aleutian). In fact, fault slip on these out-of- major growth fault (FT1) produced additional
potentially as early as the middle Eocene. This sequence thrusts is commonly synchronous with basal shear stress, which strengthened the cou-
differential evolution is commonly expressed active sedimentation in wedge-slope basins (e.g., pling of the interface between the wedge base
by temporal-spatial changes in the tectonic Barnes et al., 2010; Moore et al., 2015). This and décollement. Our results suggest that from
configuration and depositional environments. implies that sediment reworking plays a signifi- the Neogene to present day, both highly oblique
In the Burmese forearc region, for example, cant role in governing the vertical growth of the plate convergence and sediment reworking have
the Oligocene Minbu subbasin exhibits tectonic outer wedge, although it may not be the primary played an important role in controlling the verti-
and depositional settings similar to those of the factor. Detailed analyses of fault structures and cal expansion and the architecture of the outer
present-day Ayeyarwady delta within the Prome- activities, as well as the provenance of rocks wedge in the southern Indo-Burman Ranges.
Irrawaddy subbasin (Fig. 1; Gough et al., 2020; preserved in wedge slope basins, may provide
Ridd and Racey, 2015). High-resolution plate additional information on wedge deformation at ACKNOWLEDGMENTS
kinematic restorations of India, the West Burma these margins.
This work was funded by the National Natural
Terrane, Sundaland, and other relics of oceanic Science Foundation of China (no. 42002130), China
arcs within this subduction zone could help us CONCLUSIONS Postdoctoral Science Foundation (no. 2019M662742),
better understand the diachronous development and Open Research Foundation of Key Laboratory of
of convergent margins; however, plate kinematic Our results reveal a broad shelf terrace domi- Tectonics and Petroleum Resources, Ministry of Edu-
cation (no. TPR-2020-2). The authors are grateful to
restoration in the study area and adjacent regions nated by overfilled thrust-top basins, extensional Yani Najman (Lancaster University), Edward Sobel
is still heavily debated (Westerweel et al., 2019; faults, and numerous shale diapirs, as well as a (Universität Potsdam), Songling Yang (CNOOC), and
Mitchell, 2018; Morley et al., 2021). narrow deformation front characterized by high Chuanbo Shen (China University of Geosciences), for
slope angles (5.7–9.8°) and very limited fron- several conversations on topics related to this manu-
script, to Zhaochu Hu for zircon U-Pb analysis, and
Implications for the Evolution of tal accretion in the outer wedge of the southern
to Peng Deng, Qing Ye, Lulu Wu, and Shihao Hao for
Accretionary Wedges at Convergent Plate Indo-Burman Ranges. The extensional faults seismic data interpretation. We thank Paul O’Sullivan
Margins within the shelf terrace were divided into three (now at GeoSeps Services, Moscow, Idaho, USA) for
populations, which are, from oldest to youngest: measuring the apatite fission-track parameters and
In a broader context, the results of this study NW-striking late Paleogene to early Pliocene Margaret B. Donelick for collecting the LA-ICP-MS
data, both during their apatite to zircon days. Sugges-
demonstrate how an accretionary wedge with listric-style faults (Fault Set 1), NE-striking tions by Chris Morley, two anonymous reviewers, and
a wide shelf terrace and narrow deformation Neogene to present-day planar extensional faults editors W.J. Xiao, J. Waldron, and K. Cutts improved
front may develop as a consequence of highly (Fault Set 2), and WNW-striking Pliocene intra- the manuscript. P. Zhang welcomes the birth of his
oblique plate convergence and sediment rework- formational faults (Fault Set 3). The NE-striking daughter with this article.
ing of overriding plate rocks. This adds to our planar faults (Fault Set 2), which are negative
REFERENCES CITED
understanding of the mechanisms that lead to flower-like structures, are interpreted to be a
the formation of accretionary wedges in various result of dextral strike-slip tectonics. Within the Adam, J., Klaeschen, D., Kukowski, N., and Flueh, E., 2004,
tectonic contexts. deformation front, three groups of thrust faults Upward delamination of Cascadia Basin sediment in-
fill with landward frontal accretion thrusting caused by
Classically, the formation of a wedge at accre- (FT1, FT2, and FT3) nucleated “in sequence.” rapid glacial age material flux: Tectonics, v. 23, no. 3,
tionary-type margins with such geomorphologi- The oldest thrust fault (FT1) is a long-lived https://doi​.org​/10​.1029​/2002TC001475.
cal and structural characteristics is attributed growth fault with a convex-up geometry that Advokaat, E.L., Bongers, M.L.M., Rudyawan, A., BouDagh-
er-Fadel, M.K., Langereis, C.G., and van Hinsbergen,
to the occurrence of an out-of-sequence thrust governs the rapid vertical stack of the prograda- D.J.J., 2018, Early Cretaceous origin of the Woyla Arc
and/or passage of seamounts along subduction tional reflectors within the shelf terrace. (Sumatra, Indonesia) on the Australian plate: Earth and
Planetary Science Letters, v. 498, p. 348–361, https://
margins (Bangs et al., 2006; Barnes et al., 2010; The narrow deformation front and flower-like doi​.org​/10​.1016​/j​.epsl​.2018​.07​.001.
Davidson et al., 2020; Frederik et al., 2020; structures are roughly consistent with analog Aitchison, J.C., Ao, A., Bhowmik, S., Clarke, G.L., Ire-
Moore et al., 2007; Mountjoy and Barnes, 2011; modeling in the laboratory of highly oblique land, T.R., Kachovich, S., Lokho, K., Stojanovic,
D., Roeder, T., Truscott, N., Zhen, Y., and Zhou, R.,
Strasser et al., 2009). Subducting seamounts and convergence conditions, which suggests that 2019, Tectonic evolution of the western margin of the
rough seafloor topography can induce transient increasing convergence obliquity causes a hori- Burma microplate based on new fossil and radiometric

2370 Geological Society of America Bulletin, v. 135, no. 9/10

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Accretionary wedge development at oblique convergent margins

age ­constraints: Tectonics, v. 38, no. 5, p. 1718–1741, Bertrand, G., and Rangin, C., 2003, Tectonics of the western D.A., 2009, Gangdese arc detritus within the eastern
https://doi​.org​/10​.1029​/2018TC005049. margin of the Shan plateau (central Myanmar): Implica- Himalayan Neogene foreland basin: Implications for
Allen, R., Najman, Y., Carter, A., Barfod, D., Bickle, M.J., tion for the India–Indochina oblique convergence since the Neogene evolution of the Yalu–Brahmaputra River
Chapman, H.J., Garzanti, E., Vezzoli, G., Ando, S., the Oligocene: Journal of Asian Earth Sciences, v. 21, system: Earth and Planetary Science Letters, v. 285, no.
and Parrish, R.R., 2008, Provenance of the Tertiary no. 10, p. 1139–1157, https://doi​.org​/10​.1016​/S1367- 1–2, p. 150–162, https://doi.​ org/​ 10​.1016​/j​.epsl​.2009​.06​
sedimentary rocks of the Indo-Burman Ranges, Burma 9120(02)00183-9. .005.
(Myanmar): Burman arc or Himalayan-derived?: Jour- Betka, P.M., Seeber, L., Thomson, S.N., Steckler, M.S., Sin- Clark, M.K., and Royden, L.H., 2000, Topographic ooze:
nal of the Geological Society, v. 165, no. 6, p. 1045– cavage, R., and Zoramthara, C., 2018, Slip-partitioning Building the eastern margin of Tibet by lower crustal
1057, https://doi​.org​/10​.1144​/0016-76492007-143. above a shallow, weak décollement beneath the Indo- flow: Geology, v. 28, no. 8, p. 703–706, https://doi​
Arboit, F., Min, M., Chew, D., Mitchell, A., Drost, K., Baden- Burman accretionary prism: Earth and Planetary Sci- .org​/10​.1130​/0091-7613(2000)28<703:TOBTEM>2​
szki, E., and Daly, J.S., 2021, Constraining the links ence Letters, v. 503, p. 17–28, https://doi​.org​/10​.1016​ .0.CO;2.
between the Himalayan belt and the Central Myanmar /j​.epsl​.2018​.09​.003. Contreras-Reyes, E., Flueh, E.R., and Grevemeyer, I., 2010,
Basins during the Cenozoic: An integrated multi-proxy Blum, M., Rogers, K., Gleason, J., Najman, Y., Cruz, J., and Tectonic control on sediment accretion and subduction
detrital geochronology and trace-element geochemistry Fox, L., 2018, Allogenic and autogenic signals in the off south central Chile: Implications for coseismic rup-
study: Geoscience Frontiers, v. 12, no. 2, p. 657–676, stratigraphic record of the deep-sea Bengal Fan: Scien- ture processes of the 1960 and 2010 megathrust earth-
https://doi​.org​/10​.1016​/j​.gsf​.2020​.05​.024. tific Reports, v. 8, no. 1, 7973, https://doi​.org​/10​.1038​ quakes: Tectonics, v. 29, no. 6, https://doi​.org​/10​.1029​
Aung, T., 1982, Geological and geophysical map of Pyay /s41598-018-25819-5. /2010TC002734.
Embayment and Ayeyarwady Delta Area: Internal Booth-Rea, G., Klaeschen, D., Grevemeyer, I., and Reston, Copley, A., and McKenzie, D., 2007, Models of crustal flow
Report 851.M.J.N.K.O.L.P.T.M.94.A.G.C.D, scale T., 2008, Heterogeneous deformation in the Cascadia in the India-Asia collision zone: Geophysical Journal
1:250,000. convergent margin and its relation to thermal gradient International, v. 169, no. 2, p. 683–698, https://doi​.org​
Ball, T.V., Penney, C.E., Neufeld, J.A., and Copley, A.C., (Washington, NW USA): Tectonics, v. 27, no. 4, https:// /10​.1111​/j​.1365-246X​.2007​.03343​.x.
2019, Controls on the geometry and evolution of doi​.org​/10​.1029​/2007TC002209. Cruz, L., Malinski, J., Wilson, A., Take, W.A., and Hilley,
thin-skinned fold-thrust belts, and applications to the Bracciali, L., Najman, Y., Parrish, R.R., Akhter, S.H., and G., 2010, Erosional control of the kinematics and ge-
Makran accretionary prism and Indo–Burman Rang- Millar, I., 2015, The Brahmaputra tale of tectonics and ometry of fold-and-thrust belts imaged in a physical and
es: Geophysical Journal International, v. 218, no. 1, erosion: Early Miocene river capture in the Eastern numerical sandbox: Journal of Geophysical Research:
p. 247–267, https://doi​.org​/10​.1093​/gji​/ggz139. Himalaya: Earth and Planetary Science Letters, v. 415, Solid Earth, v. 115, no. B09404, https://doi.​ org/​ 10.​ 1029​
Bandopadhyay, P.C., van Hinsbergen, D.J.J., Bandyopadhy- p. 25–37, https://doi​.org​/10​.1016​/j​.epsl​.2015​.01​.022. /2010JB007472.
ay, D., Licht, A., Advokaat, E.L., Plunder, A., Ghosh, Braun, J., and Beaumont, C., 1995, Three-dimensional nu- Cruz, L., Malinski, J., Hernandez, M., Take, A., and Hilley,
B., Dasgupta, A., and Trabucho-Alexandre, J.P., 2022, merical experiments of strain partitioning at oblique G., 2011, Erosional control of the kinematics of the
Paleogeography of the West Burma Block and the plate boundaries: Implications for contrasting tec- Aconcagua fold-and-thrust belt from numerical simu-
eastern Neotethys Ocean: Constraints from Cenozoic tonic styles in the southern Coast Ranges, California, lations and physical experiments: Geology, v. 39, no. 5,
sediments shed onto the Andaman-Nicobar ophiolites: and central South Island, New Zealand: Journal of p. 439–442, https://doi​.org​/10​.1130​/G31675​.1.
Gondwana Research, v. 103, p. 335–361, https://doi.​ org​ Geophysical Research: Solid Earth, v. 100, no. B9, Curray, J.R., 1991, Possible greenschist metamorphism at
/10​.1016​/j​.gr​.2021​.10​.011. p. 18,059–18,074, https://doi​.org​/10​.1029​/95JB01683. the base of a 22-km sedimentary section, Bay of Ben-
Bangs, N., Westbrook, G.K., Ladd, J.W., and Buhl, P., 1990, Brunnschweiler, R.O., 1966, On the geology of the Indo- gal: Geology, v. 19, no. 11, p. 1097–1100, https://doi​
Seismic velocities from the Barbados Ridge Complex: burman ranges: Journal of the Geological Society of .org​/10​.1130​/0091-7613(1991)019<1097:PGMATB>2​
Indicators of high pore fluid pressures in an accretion- Australia, v. 13, no. 1, p. 137–194, https://doi​.org​/10​ .3.CO;2.
ary complex: Journal of Geophysical Research: Solid .1080​/00167616608728608. Curray, J.R., 2005, Tectonics and history of the Andaman
Earth, v. 95, no. B6, p. 8767–8782, https://doi​.org​/10​ Burbidge, D.R., and Braun, J., 1998, Analogue models of Sea region: Journal of Asian Earth Sciences, v. 25, no.
.1029​/JB095iB06p08767. obliquely convergent continental plate boundaries: 1, p. 187–232, https://doi​.org​/10​.1016​/j​.jseaes​.2004​.09​
Bangs, N.L., Shipley, T.H., Gulick, S., Moore, G.F., Kuromo- Journal of Geophysical Research: Solid Earth, v. 103, .001.
to, S., and Nakamura, Y., 2004, Evolution of the Nankai no. B7, p. 15,221–15,237, https://doi​.org​/10​.1029​ Dahlen, F.A., Suppe, J., and Davis, D., 1984, Mechanics of
Trough décollement from the trench into the seismo- /98JB00751. fold-and-thrust belts and accretionary wedges; cohesive
genic zone: Inferences from three-dimensional seismic Bürgi, P., Hubbard, J., Akhter, S.H., and Peterson, D.E., Coulomb theory: Journal of Geophysical Research:
reflection imaging: Geology, v. 32, no. 4, p. 273–276, 2021, Geometry of the décollement below eastern Ban- Solid Earth, v. 89, no. B12, p. 10,087–10,101, https://
https://doi​.org​/10​.1130​/G20211​.2. gladesh and implications for seismic hazard: Journal doi​.org​/10​.1029​/JB089iB12p10087.
Bangs, N.L.B., Gulick, S.P.S., and Shipley, T.H., 2006, Sea- of Geophysical Research: Solid Earth, v. 126, no. 8, Davidson, S.R., Barnes, P.M., Pettinga, J.R., Nicol, A.,
mount subduction erosion in the Nankai Trough and its https://doi​.org​/10​.1029​/2020JB021519. Mountjoy, J.J., and Henrys, S.A., 2020, Conjugate
potential impact on the seismogenic zone: Geology, v. 34, Burma Earth Sciences Research Division, 1977, Geological strike-slip faulting across a subduction front driven by
no. 8, p. 701–704, https://doi​.org​/10​.1130​/G22451​.1. map of the Socialist Republic of the Union of Burma: incipient seamount subduction: Geology, v. 48, no. 5,
Bannert, D., Lyen, A.S., and Htay, T., 2011, The Geology of Burma, Security Printing Works, scale 1:1,000,000. p. 493–498, https://doi​.org​/10​.1130​/G47154​.1.
Indoburman Ranges in Myanmar: Geologisches Jahr- Burnham, A.K., and Sweeney, J.J., 1989, A chemical kinetic Davis, D., Suppe, J., and Dahlen, F.A., 1983, Mechanics
burch Reihe B, p. 101. model of vitrinite maturation and reflectance: Geochim- of fold- and -thrust belts and accretionary wedges:
Bao, C., Chen, Y.L., Bu, X.F., Chen, X., and Li, D.P., 2015, ica et Cosmochimica Acta, v. 53, p. 2649–2657, https:// Journal of Geophysical Research: Solid Earth, v. 88,
LA-ICP-MS U-Pb ages and Hf isotopic compositions doi​.org​/10​.1016​/0016-7037(89)90136-1. no. NB2, p. 1153–1172, https://doi​. org​/ 10​. 1029​
of detrital zircons in the sediments of the Nujiang River, Burrett, C., Zaw, K., Meffre, S., Lai, C.K., Khositanont, S., /JB088iB02p01153.
Yunnan Province, and their geological significance [in Chaodumrong, P., Udchachon, M., Ekins, S., and Hal- Dickinson, W.R., and Gehrels, G.E., 2009, Use of U–Pb ages
Chinese with English abstract]: Geological Bulletin of pin, J., 2014, The configuration of Greater Gondwa- of detrital zircons to infer maximum depositional ages
China, v. 34, no. 8, p. 1413–1425. na—Evidence from LA ICPMS, U-Pb geochronology of strata: A test against a Colorado Plateau Mesozoic
Barber, A.J., and Crow, M.J., 2009, Structure of Sumatra and of detrital zircons from the Palaeozoic and Mesozoic database: Earth and Planetary Science Letters, v. 288,
its implications for the tectonic assembly of Southeast of Southeast Asia and China: Gondwana Research, no. 1–2, p. 115–125, https://doi.​ org/​ 10.​ 1016/​ j.​ epsl.​ 2009​
Asia and the destruction of Paleotethys: The Island Arc, v. 26, no. 1, p. 31–51, https://doi​.org​/10​.1016​/j​.gr​.2013​ .09​.013.
v. 18, no. 1, p. 3–20, https://doi​.org​/10​.1111​/j​.1440- .05​.020. Ding, L., Goswami, T.K., Cai, F., Baral, U., Sarmah, R.K.,
1738​.2008​.00631​.x. Cai, F., Ding, L., Laskowski, A.K., Kapp, P., Wang, H., Xu, and Bezbaruah, D., 2022, Detrital zircon U-Pb ages of
Barnes, P.M., and de Lépinay, B.M., 1997, Rates and me- Q., and Zhang, L., 2016, Late Triassic paleogeographic Tertiary sequences (Palaeocene–Miocene): Inner Fold
chanics of rapid frontal accretion along the very reconstruction along the Neo–Tethyan Ocean margins, Belt and Belt of Schuppen, Indo-Myanmar Ranges,
obliquely convergent southern Hikurangi margin, New southern Tibet: Earth and Planetary Science Letters, India: Geological Journal, in press, https://doi​.org​/10​
Zealand: Journal of Geophysical Research: Solid Earth, v. 435, p. 105–114, https://doi​.org​/10​.1016​/j​.epsl​.2015​ .1002​/gj​.4446.
v. 102, no. B11, p. 24,931–24,952, https://doi​.org​/10​ .12​.027. Fan, S., Ding, L., Murphy, M.A., Yao, W., and Yin, A., 2017,
.1029​/97JB01384. Cai, F., Ding, L., Yao, W., Laskowski, A.K., Xu, Q., Zhang, J., Late Paleozoic and Mesozoic evolution of the Lhasa
Barnes, P.M., Davy, B.W., Sutherland, R., and Delteil, J., and Sein, K., 2017, Provenance and tectonic evolution of Terrane in the Xainza area of southern Tibet: Tectono-
2002, Frontal accretion and thrust wedge evolution un- Lower Paleozoic–Upper Mesozoic strata from Sibumasu physics, v. 721, p. 415–434, https://doi​.org​/10​.1016​/j​
der very oblique plate convergence; Fiordland Basin, Terrane, Myanmar: Gondwana Research, v. 41, p. 325– .tecto​.2017​.10​.022.
New Zealand: Basin Research, v. 14, no. 4, p. 439–466, 336, https://doi​.org​/10​.1016​/j​.gr​.2015​.03​.005. Fillon, C., Huismans, R.S., and van der Beek, P., 2013, Syn-
https://doi​.org​/10​.1046​/j​.1365-2117​.2002​.00178​.x. Cai, F., Ding, L., Zhang, Q., Orme, D. A., Wei, H., Li, J., tectonic sedimentation effects on the growth of fold and
Barnes, P.M., Lamarche, G., Bialas, J., Henrys, S., Pecher, Zhang, J., Zaw, T., and Sein, K., 2020, Initiation and thrust belts: Geology, v. 41, p. 83–86, https://doi​.org​/10​
I., Netzeband, G.L., Greinert, J., Mountjoy, J.J., Ped- evolution of forearc basins in the Central Myanmar .1130​/G33531​.1.
ley, K., and Crutchley, G., 2010, Tectonic and geologi- Depression: Geological Society of America Bulletin, Fitch, T., 1972, Plate convergence, transcurrent faults, and
cal framework for gas hydrates and cold seeps on the v. 132, no. 5–6, p. 1066–1082, https://doi​.org​/10​.1130​ internal deformation adjacent to Southeast Asia and
Hikurangi subduction margin, New Zealand: Marine /B35301​.1. the Western Pacific: Journal of Geophysical Research,
Geology, v. 272, no. 1–4, p. 26–48, https://doi​.org​/10​ Cina, S.E., Yin, A., Grove, M., Dubey, C.S., Shukla, D.P., v. 77, no. 23, p. 4432–4460, https://doi​. org​/10​. 1029​
.1016​/j​.margeo​.2009​.03​.012. Lovera, O.M., Kelty, T.K., Gehrels, G.E., and Foster, /JB077i023p04432.

Geological Society of America Bulletin, v. 135, no. 9/10 2371

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Peng Zhang et al.

Frederik, M.C.G., Gulick, S.P.S., and Miller, J.J., 2020, Ef- Earth and Planetary Science Letters, v. 397, p. 145–158, Caribbean margin: Constraints from 2-D seismic re-
fect on subduction of deeply buried seamounts offshore https://doi​.org​/10​.1016​/j​.epsl​.2014​.04​.026. flection data and potential fields interpretation, in Lal-
of Kodiak Island: Tectonics, v. 39, no. 7, https://doi​.org​ Leever, K.A., Gabrielsen, R.H., Sokoutis, D., and Willing- lemand, S., and Funiciello, F., eds., Subduction Zone
/10​.1029​/2019TC005710. shofer, E., 2011, The effect of convergence angle on the Geodynamics: Frontiers in Earth Sciences: Berlin,
Frisch, W., Meschede, M., and Blakey, R., 2011, Plate kinematic evolution of strain partitioning in transpres- Springer, p. 247–272, https://doi​.org​/10​.1007​/978-3-
Tectonics: Continental Drift and Mountain Building: sional brittle wedges: Insight from analog modeling and 540-87974-9_13.
Berlin, Springer, 212 p., https://doi​.org​/10​.1007​/978- high-resolution digital image analysis: Tectonics, v. 30, Maurin, T., and Rangin, C., 2009a, Structure and kinematics
3-540-76504-2. no. 2, https://doi​.org​/10​.1029​/2010TC002823. of the Indo-Burmese Wedge: Recent and fast growth of
Garzanti, E., Wang, J., Vezzoli, G., and Limonta, M., 2016, Lei, Q.P., Wu, X., and Wan, X.Q., 2009, The Eocene–Plio- the outer wedge: Tectonics, v. 28, no. 2, https://doi​.org​
Tracing provenance and sediment fluxes in the Ir- cene planktonic foraminifera from Ramree Island, /10​.1029​/2008TC002276.
rawaddy River basin (Myanmar): Chemical Geology, Burma [in Chinese with English abstract]: Acta Micro- Maurin, T., and Rangin, C., 2009b, Impact of the 90°E ridge
v. 440, p. 73–90, https://doi​.org​/10​.1016​/j​.chemgeo​ palaeontologica Sinica, v. 26, no. 4, p. 323–330. at the Indo-Burmese subduction zone imaged from deep
.2016​.06​.010. Leier, A.L., Kapp, P., Gehrels, G.E., and DeCelles, seismic reflection data: Marine Geology, v. 266, p. 143–
Gehrels, G., Kapp, P., DeCelles, P., Pullen, A., Blakey, R., P.G., 2007, Detrital zircon geochronology of 155, https://doi​.org​/10​.1016​/j​.margeo​.2009​.07​.015.
Weislogel, A., Ding, L., Guynn, J., Martin, A., McQuar- Carboniferous?Cretaceous strata in the Lhasa terrane, Maurin, T., Masson, F., Rangin, C., Min, U.T., and Col-
rie, N., and Yin, A., 2011, Detrital zircon geochronol- Southern Tibet: Basin Research, v. 19, no. 3, p. 361– lard, P., 2010, First Global Positioning System results
ogy of pre-Tertiary strata in the Tibetan-Himalayan 378, https://doi​ . org​ / 10​ . 1111​ / j​ . 1365-2117​ . 2007​ in northern Myanmar; constant and localized slip rate
orogen: Tectonics, v. 30, no. 5, https://doi​.org​/10​.1029​ .00330​.x. along the Sagaing Fault: Geology, v. 38, no. 7, p. 591–
/2011TC002868. Li, D., Chen, Y., Hou, K., and Luo, Z., 2016, Origin and evo- 594, https://doi​.org​/10​.1130​/G30872​.1.
Gopala Rao, D., Krishna, K.S., and Sar, D., 1997, Crustal lution of the Tengchong block, southeastern margin of McCaffrey, R., 1992, Oblique plate convergence, slip vec-
evolution and sedimentation history of the Bay of Ben- the Tibetan Plateau: Zircon U–Pb and Lu–Hf isotopic tors, and forearc deformation: Journal of Geophysical
gal since the Cretaceous: Journal of Geophysical Re- evidence from the (meta-) sedimentary rocks and intru- Research: Solid Earth, v. 97, no. B6, p. 8905–8915,
search: Solid Earth, v. 102, no. B8, p. 17,747–17,768, sions: Tectonophysics, v. 687, p. 245–256, https://doi​ https://doi​.org​/10​.1029​/92JB00483.
https://doi​.org​/10​.1029​/96JB01339. .org​/10​.1016​/j​.tecto​.2016​.07​.008. McClay, K.R., Whitehouse, P.S., Dooley, T., and Richards,
Gough, A., Hall, R., and BouDagher-Fadel, M.K., 2020, Licht, A., France-Lanord, C., Reisberg, L., Fontaine, C., Soe, M., 2004, 3-D evolution of fold and thrust belts formed
Mid-Cenozoic fluvio-deltaic to marine environments A.N., and Jaeger, J.J., 2013, A palaeo Tibet–Myanmar by oblique convergence: Marine and Petroleum Geol-
of the Salin Sub-basin, Central Myanmar: Journal of connection? Reconstructing the late Eocene drainage ogy, v. 21, no. 7, p. 857–877, https://doi​.org​/10​.1016​/j​
Asian Earth Sciences, v. 190, https://doi​.org​/10​.1016​/j​ system of central Myanmar using a multi-proxy ap- .marpetgeo​.2004​.03​.009.
.jseaes​.2019​.104143. proach: Journal of the Geological Society, v. 170, no. McNeill, L.C., and Henstock, T.J., 2014, Forearc structure
Gramann, F., 1974, Some palaeontological data on the Trias- 6, p. 929–939, https://doi​.org​/10​.1144​/jgs2012-126. and morphology along the Sumatra-Andaman Subduc-
sic and Cretaceous of the western part of Burma (Ara- Licht, A., Dupont-Nivet, G., Win, Z., Swe, H.H., Kaythi, tion Zone: Tectonics, v. 33, no. 2, p. 112–134, https://
kan Islands, Arakan Yoma, western outcrops of Central M., Roperch, P., Ugrai, T., Littell, V., Park, D., Wester- doi​.org​/10​.1002​/2012TC003264.
Basin): Newsletters on Stratigraphy, v. 3, p. 277–290, weel, J., Jones, D., Poblete, F., Aung, D.W., Huang, H., McNeill, L.C., Dugan, B., Backman, J., Pickering, K.T.,
https://doi​.org​/10​.1127​/nos​/3/1974​/277. Hoorn, C., and Sein, K., 2019, Paleogene evolution of Pouderoux, H.F.A., Henstock, T.J., Petronotis, K.E.,
Grando, G., and McClay, K., 2007, Morphotectonics do- the Burmese forearc basin and implications for the his- Carter, A., Chemale, F., Milliken, K.L., Kutterolf, S.,
mains and structural styles in the Makran accretionary tory of India-Asia convergence: Geological Society of Mukoyoshi, H., Chen, W., Kachovich, S., Mitchison,
prism, offshore Iran: Sedimentary Geology, v. 196, no. America Bulletin, v. 131, no. 5–6, p. 730–748, https:// F.L., Bourlange, S., Colson, T.A., Frederik, M.C.G.,
1–4, p. 157–179, https://doi​.org​/10​.1016​/j​.sedgeo​.2006​ doi​.org​/10​.1130​/B35002​.1. Guèrin, G., Hamahashi, M., House, B.M., Hüpers, A.,
.05​.030. Licht, A., Win, Z., Westerweel, J., Cogné, N., Morley, C.K., Jeppson, T.N., Kenigsberg, A.R., Kuranaga, M., Nair,
Hoke, G.D., Liu-Zeng, J., Hren, M.T., Wissink, G.K., and Chantraprasert, S., Poblete, F., Ugrai, T., Nelson, B., N., Owari, S., Shan, Y., Song, I., Torres, M.E., Vannuc-
Garzione, C.N., 2014, Stable isotopes reveal high Aung, D.W., and Dupont-Nivet, G., 2020, Magmatic his- chi, P., Vrolijk, P.J., Yang, T., Zhao, X., and Thomas,
southeast Tibetan Plateau margin since the Paleogene: tory of central Myanmar and implications for the evolu- E., 2017, Understanding Himalayan erosion and the
Earth and Planetary Science Letters, v. 394, p. 270–278, tion of the Burma Terrane: Gondwana Research, v. 87, significance of the Nicobar Fan: Earth and Planetary
https://doi​.org​/10​.1016​/j​.epsl​.2014​.03​.007. p. 303–319, https://doi​.org​/10​.1016​/j​.gr​.2020​.06​.016. Science Letters, v. 475, p. 134–142, https://doi​.org​/10​
Jain, M., Das, P.S., and Bandyopadhyay, B., 2010, Structural Limonta, M., Resentini, A., Carter, A., Bandopadhyay, P.C., .1016​/j​.epsl​.2017​.07​.019.
framework and deep-marine depositional environments and Garzanti, E., 2017, Provenance of Oligocene Anda- McQuarrie, N., Robinson, D., Long, S., Tobgay, T., Grujic,
of Miocene–Pleistocene sequence in western offshore man sandstones (Andaman–Nicobar Islands): Ganga– D., Gehrels, G., and Ducea, M., 2008, Preliminary
Myanmar: Hyderabad, The Society of Petroleum Geo- Brahmaputra or Irrawaddy derived?, in Bandopadhyay, stratigraphic and structural architecture of Bhutan:
physicists India, 8th Biennial International Conference P.C.C.A., ed., The Andaman–Nicobar Accretionary Implications for the along strike architecture of the Hi-
& Exposition on Petroleum Geophysics, P-58, p. 1–8. Ridge: Geology, Tectonics and Hazards: Geological malayan system: Earth and Planetary Science Letters,
Jonell, T.N., Giosan, L., Clift, P.D., Carter, A., Bretschneider, Society, London, Memoir 47, p. 141–152, https://doi​ v. 272, no. 1–2, p. 105–117, https://doi​.org​/10​.1016​/j​
L., Hathorne, E.C., Barbarano, M., Garzanti, E., Vez- .org​/10​.1144​/M47​.10. .epsl​.2008​.04​.030.
zoli, G., and Naing, T., 2022, No modern Irrawaddy Lin, T.H., Chung, S.L., Tang, J.T., and Oo, T., 2017, The de- McQuarrie, N., Long, S.P., Tobgay, T., Nesbit, J.N., Gehrels,
River until the late Miocene–Pliocene: Earth and Plan- limitation between the mature and juvenile crustal prov- G., and Ducea, M.N., 2013, Documenting basin scale,
etary Science Letters, v. 584, https://doi​.org​/10​.1016​/j​ inces in SE Asia: Insights from detrital zircon U-Pb and geometry and provenance through detrital geochemical
.epsl​.2022​.117516. Hf isotopic data for the Salween drainage, Myanmar: data: Lessons from the Neoproterozoic to Ordovician
Karig, D.E., Lawrence, M.B., Moore, G.F., and Curray, J.R., Journal of Asian Earth Sciences, v. 145, p. 641–651, Lesser, Greater, and Tethyan Himalayan strata of Bhu-
1980, Structural frame work of the fore-arc basin, NW https://doi​.org​/10​.1016​/j​.jseaes​.2017​.06​.018. tan: Gondwana Research, v. 23, no. 4, p. 1491–1510,
Sumatra: Journal of the Geological Society, v. 137, no. Liu, Q.Y., van der Hilst, R.D., Li, Y., Yao, H.J., Chen, J.H., https://doi​.org​/10​.1016​/j​.gr​.2012​.09​.002.
1, p. 77–91, https://doi​.org​/10​.1144​/gsjgs​.137​.1.0077. Guo, B., Qi, S.H., Wang, J., Huang, H., and Li, S.C., Misawa, A., Hirata, K., Seeber, L., Arai, K., Nakamura, Y.,
Ketcham, R.A., Donelick, R.A., and Donelick, M.B., 2000, 2014, Eastward expansion of the Tibetan Plateau by Rahardiawan, R., Udrekh, Fujiwara, T., Kinoshita,
AFTSolve: A program for multi-kinetic modeling of crustal flow and strain partitioning across faults: Na- M., Baba, H., Kameo, K., Adachi, K., Sarukawa, H.,
apatite fission-track data: The American Mineralogist, ture Geoscience, v. 7, no. 5, p. 361–365, https://doi​.org​ Tokuyama, H., Permana, H., Djajadihardja, Y.S., and
v. 2, no. 5–6, p. 929. /10​.1038​/ngeo2130. Ashi, J., 2014, Geological structure of the offshore
Khin, K., and Myitta, 1999, Marine transgression and regres- Lüschen, E., Müller, C., Kopp, H., Engels, M., Lutz, R., Sumatra forearc region estimated from high-resolution
sion in Miocene sequences of northern Pegu (Bago) Planert, L., Shulgin, A., and Djajadihardja, Y.S., 2011, MCS reflection survey: Earth and Planetary Science
Yoma, central Myanmar: Journal of Asian Earth Sci- Structure, evolution, and tectonic activity of the east- Letters, v. 386, p. 41–51, https://doi​.org​/10​.1016​/j​.epsl​
ences, v. 17, no. 3, p. 369–393, https://doi​.org​/10​.1016​ ern Sunda forearc, Indonesia, from marine seismic in- .2013​.10​.031.
/S0743-9547(98)00065-8. vestigations: Tectonophysics, v. 508, no. 1–4, p. 6–21, Mitchell, A.H.G., 1993, Cretaceous–Cenozoic tectonic events
Khin, K., Moe, A., and Myint, M., 2020, Geology, structure https://doi​.org​/10​.1016​/j​.tecto​.2010​.06​.008. in the western Myanmar (Burma)–Assam region: Journal
and lithostratigraphic framework of the Rakhine Coast- Mallick, R., Hubbard, J.A., Lindsey, E.O., Bradley, K.E., of the Geological Society, v. 150, no. 6, p. 1089–1102,
al Ranges in western Myanmar: Implications for the Moore, J.D.P., Ahsan, A., Khorshed Alam, A.K.M., and https://doi​.org​/10​.1144​/gsjgs​.150​.6.1089.
collision of the India Plate and West Myanmar Block: Hill, E.M., 2020, Subduction initiation and the rise of Mitchell, A., 2018, Geological Belts, Plate Boundaries, and
Journal of Asian Earth Sciences, v. 196, https://doi​.org​ the Shillong Plateau: Earth and Planetary Science Let- Mineral Deposits in Myanmar: Amsterdam, Elsevier
/10​.1016​/j​.jseaes​.2020​.104332. ters, v. 543, https://doi.​ org/​ 10.​ 1016/​ j.​ epsl.​ 2020.​ 116351. Inc., 352 p.
Koons, P.O., 1994, Three-dimensional critical wedges: Tec- Mannu, U., Ueda, K., Willett, S.D., Gerya, T.V., and Stras- Moeremans, R., Singh, S.C., Mukti, M., McArdle, J., and
tonics and topography in oblique collisional orogens: ser, M., 2016, Impact of sedimentation on evolution Johansen, K., 2014, Seismic images of structural varia-
Journal of Geophysical Research: Solid Earth, v. 99, of accretionary wedges: Insights from high-resolution tions along the deformation front of the Andaman–
no. B6, p. 12,301–12,315, https://doi​.org​/10​.1029​ thermomechanical modeling: Tectonics, v. 35, no. 12, Sumatra Subduction Zone: Implications for rupture
/94JB00611. p. 2828–2846, https://doi​.org​/10​.1002​/2016TC004239. propagation and tsunamigenesis: Earth and Planetary
Lang, K.A., and Huntington, K.W., 2014, Antecedence of the Mantilla-Pimiento, A.M., Jentzsch, G., Kley, J., and Alfon- Science Letters, v. 386, p. 75–85, https://doi​.org​/10​
Yarlung–Siang–Brahmaputra River, eastern Himalaya: so-Pava, C., 2009, Configuration of the Colombian .1016​/j​.epsl​.2013​.11​.003.

2372 Geological Society of America Bulletin, v. 135, no. 9/10

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Accretionary wedge development at oblique convergent margins

Moore, G.F., Bangs, N.L., Taira, A., Kuramoto, S., Pangborn, Hydrocarbon Exploration and Production: Geological Racey, A., and Ridd, M.F., 2015, Petroleum geology of the
E., and Tobin, H.J., 2007, Three-dimensional splay fault Society, London, Special Publication 386, p. 195–216, Moattama Region, Myanmar, in Racey, A., and Ridd,
geometry and implications for tsunami generation: Sci- https://doi​.org​/10​.1144​/SP386​.10. M.F., eds., Petroleum Geology of Myanmar: Geological
ence, v. 318, no. 5853, p. 1128–1131, https://doi​.org​/10​ Najman, Y., Bickle, M., BouDagher-Fadel, M., Carter, A., Society, London, Memoir 45, p. 63–81, https://doi​.org​
.1126​/science​.1147195. Garzanti, E., Paul, M., Wijbrans, J., Willett, E., Oliver, /10​.1144​/M45​.07.
Moore, G.F., Boston, B.B., Strasser, M., Underwood, M.B., G., Parrish, R., Akhter, S.H., Allen, R., Ando, S., Chisty, Rahman, M.J.J., Xiao, W., Hossain, M.S., Yeasmin, R., Say-
and Ratliff, R.A., 2015, Evolution of tectono-sedimen- E., Reisberg, L., and Vezzoli, G., 2008, The Paleogene em, A.S.M., Ao, S., Yang, L., Abdullah, R., and Dina,
tary systems in the Kumano Basin, Nankai Trough record of Himalayan erosion: Bengal Basin, Bangla- N.T., 2020, Geochemistry and detrital zircon U–Pb
forearc: Marine and Petroleum Geology, v. 67, p. 604– desh: Earth and Planetary Science Letters, v. 273, no. dating of Pliocene–Pleistocene sandstones of the Chit-
616, https://doi​.org​/10​.1016​/j​.marpetgeo​.2015​.05​.032. 1–2, p. 1–14, https://doi​.org​/10​.1016​/j​.epsl​.2008​.04​ tagong Tripura Fold Belt (Bangladesh): Implications
Moore, G.F., Aung, L.T., Fukuchi, R., Sample, J.C., Hell- .028. for provenance: Gondwana Research, v. 78, p. 278–
ebrand, E., Kopf, A., Naing, W., Than, W.M., and Tun, Najman, Y., Allen, R., Willett, E.A.F., Carter, A., Barfod, D., 290, https://doi​.org​/10​.1016​/j​.gr​.2019​.07​.018.
T.N., 2019, Tectonic, diapiric and sedimentary chaotic Garzanti, E., Wijbrans, J., Bickle, M.J., Vezzoli, G., Rangin, C., 2017, Active and recent tectonics of the Burma
rocks of the Rakhine coast, western Myanmar: Gond- Ando, S., Oliver, G., and Uddin, M.J., 2012, The re- Platelet in Myanmar, in Barber, A.J., Zaw, K., and
wana Research, v. 74, p. 126–143, https://doi​.org​/10​ cord of Himalayan erosion preserved in the sedimentary Crow, M.J., eds., Myanmar: Geology, Resources and
.1016​/j​.gr​.2019​.04​.006. rocks of the Hatia Trough of the Bengal Basin and the Tectonics: Geological Society, London, Memoir 48,
Morgan, J.K., and Bangs, N.L., 2017, Recognizing seamount- Chittagong Hill Tracts, Bangladesh: Basin Research, p. 53–64, https://doi​.org​/10​.1144​/M48​.3.
forearc collisions at accretionary margins; insights from v. 24, no. 5, p. 499–519, https://doi.​ org/​ 10.​ 1111/​ j.​ 1365- Rangin, C., and Sibuet, J., 2017, Structure of the northern
discrete numerical simulations: Geology, v. 45, no. 7, 2117​.2011​.00540​.x. Bay of Bengal offshore Bangladesh: Evidences from
p. 635–638, https://doi​.org​/10​.1130​/G38923​.1. Najman, Y., Sobel, E.R., Millar, I., Stockli, D.F., Govin, new multi-channel seismic data: Marine and Petroleum
Morley, C.K., 2007, Interaction between critical wedge ge- G., Lisker, F., Garzanti, E., Limonta, M., Vezzoli, G., Geology, v. 84, p. 64–75, https://doi​.org​/10​.1016​/j​
ometry and sediment supply in a deep-water fold belt: Copley, A., Zhang, P., Szymanski, E., and Kahn, A., .marpetgeo​.2017​.03​.020.
Geology, v. 35, no. 2, p. 139–142, https://doi​.org​/10​ 2020, The exhumation of the Indo-Burman Ranges, Rangin, C., Maurin, T., and Masson, F., 2013, Combined ef-
.1130​/G22921A​.1. Myanmar: Earth and Planetary Science Letters, v. 530, fects of Eurasia/Sunda oblique convergence and East-
Morley, C.K., 2017, Syn-kinematic sedimentation at a re- https://doi​.org​/10​.1016​/j​.epsl​.2019​.115948. Tibetan crustal flow on the active tectonics of Burma:
leasing splay in the northern Minwun Ranges, Sagaing Najman, Y., Sobel, E.R., Millar, I., Luan, X., Zapata, S., Journal of Asian Earth Sciences, v. 76, p. 185–194,
Fault zone, Myanmar: Significance for fault timing Garzanti, E., Parra, M., Vezzoli, G., Zhang, P., Wa https://doi​.org​/10​.1016​/j​.jseaes​.2013​.05​.018.
and displacement: Basin Research, v. 29, p. 684–700, Aung, D., Paw, S.M.T.L., and Lwin, T.N., 2022, The Reiners, P.W., and Brandon, M.T., 2006, Using thermochro-
https://doi​.org​/10​.1111​/bre​.12201. timing of collision between Asia and the West Burma nology to understand orogenic erosion: Annual Review
Morley, C.K., and Alvey, A., 2015, Is spreading prolonged, Terrane, and the development of the Indo-Burman of Earth and Planetary Sciences, v. 34, p. 419–466,
episodic or incipient in the Andaman Sea? Evidence Ranges: Tectonics, v. 41, no. 7, https://doi​.org​/10​.1029​ https://doi​ . org​ / 10​ . 1146​ / annurev​ . earth​ . 34​ . 031405​
from deepwater sedimentation: Journal of Asian Earth /2021TC007057. .125202.
Sciences, v. 98, p. 446–456, https://doi​.org​/10​.1016​/j​ Nicol, A., Mazengarb, C., Chanier, F., Rait, G., Uruski, C., Replumaz, A., Negredo, A.M., Guillot, S., and Villaseñor,
.jseaes​.2014​.11​.033. and Wallace, L., 2007, Tectonic evolution of the active A., 2010, Multiple episodes of continental subduction
Morley, C.K., and Arboit, F., 2019, Dating the onset of mo- Hikurangi subduction margin, New Zealand, since the during India/Asia convergence: Insight from seismic
tion on the Sagaing fault: Evidence from detrital zircon Oligocene: Tectonics, v. 26, no. 4, https://doi​.org​/10​ tomography and tectonic reconstruction: Tectonophys-
and titanite U-Pb geochronology from the North Min- .1029​/2006TC002090. ics, v. 483, no. 1–2, p. 125–134, https://doi​.org​/10​.1016​
wun Basin, Myanmar: Geology, v. 47, no. 6, p. 581– Nielsen, C., Chamot-Rooke, N., and Rangin, C., 2004, /j​.tecto​.2009​.10​.007.
585, https://doi​.org​/10​.1130​/G46321​.1. From partial to full strain partitioning along the Indo- Ridd, M.F., and Racey, A., 2015, Onshore petroleum geology
Morley, C.K., Naing, T.T., Searle, M., and Robinson, S.A., Burmese hyper-oblique subduction: Marine Geology, of Myanmar: Central Burma Depression, in Racey, A.,
2020, Structural and tectonic development of the Indo- v. 209, no. 1–4, p. 303–327, https://doi​.org​/10​.1016​/j​ and Ridd, M.F., eds., Petroleum Geology of Myanmar:
Burma Ranges: Earth-Science Reviews, v. 200, https:// .margeo​.2004​.05​.001. Geological Society, London, Memoir 45, p. 21–50,
doi​.org​/10​.1016​/j​.earscirev​.2019​.102992. Nielsen, S.B., Clausen, O.R., and McGregor, E., 2017, https://doi​.org​/10​.1144​/M45​.04.
Morley, C.K., Chantraprasert, S., Kongchum, J., and Chenoll, Basin%Ro: A vitrinite reflectance model derived from Robinson, R.A.J., Brezina, C.A., Parrish, R.R., Horstwood,
K., 2021, The West Burma Terrane, a review of recent basin and laboratory data: Basin Research, v. 29, no. 1, M.S.A., Oo, N.W., Bird, M.I., Thein, M., Walters, A.S.,
paleo-latitude data, its geological implications and con- p. 515–536, https://doi​.org​/10​.1111​/bre​.12160. Oliver, G.J.H., and Zaw, K., 2014, Large rivers and oro-
straints: Earth-Science Reviews, v. 220, https://doi​.org​ Noda, A., 2016, Forearc basins: Types, geometries, and rela- gens: The evolution of the Yarlung Tsangpo-Irrawaddy
/10​.1016​/j​.earscirev​.2021​.103722. tionships to subduction zone dynamics: Geological So- system and the eastern Himalayan Syntaxis: Gondwana
Mortera-Gutiérrez, C.A., Scholl, D.W., and Carlson, R.L., ciety of America Bulletin, v. 128, no. 5–6, p. 879–895, Research, v. 26, p. 112–121, https://doi​.org​/10​.1016​/j​
2003, Fault trends on the seaward slope of the Aleutian https://doi​.org​/10​.1130​/B31345​.1. .gr​.2013​.07​.002.
Trench: Implications for a laterally changing stress field Noda, A., Koge, H., Yamada, Y., Miyakawa, A., and Ashi, J., Ruh, J.B., 2016, Submarine landslides caused by seamounts
tied to a westward increase in oblique convergence: 2020, Forearc basin stratigraphy resulting from syntec- entering accretionary wedge systems: Terra Nova, v. 28,
Journal of Geophysical Research: Solid Earth, v. 108, tonic sedimentation during accretionary wedge growth: no. 3, p. 163–170, https://doi​.org​/10​.1111​/ter​.12204.
no. B10, https://doi​.org​/10​.1029​/2001JB001433. Insights from sandbox analog experiments: Tectonics, Schenk, O., Peters, K.E., and Burnham, A.K., 2017, Evalu-
Mountjoy, J.J., and Barnes, P.M., 2011, Active upper plate v. 39, no. 3, https://doi​.org​/10​.1029​/2019TC006033. ation of alternatives to Easy%Ro for calibration of ba-
thrust faulting in regions of low plate interface cou- Pesicek, J.D., Thurber, C.H., Zhang, H., DeShon, H.R., sin and petroleum system models [extended abstract]:
pling, repeated slow slip events, and coastal uplift: Engdahl, E.R., and Widiyantoro, S., 2010, Teleseis- Paris, France, EAGE.
Example from the Hikurangi Margin, New Zealand: mic double-difference relocation of earthquakes along Sevastjanova, I., Hall, R., Rittner, M., Paw, S.M.T.L., Naing,
Geochemistry, Geophysics, Geosystems, v. 12, no. 1, the Sumatra-Andaman Subduction Zone using a 3-D T.T., Alderton, D.H., and Comfort, G., 2016, Myanmar
https://doi​.org​/10​.1029​/2010GC003326. model: Journal of Geophysical Research: Solid Earth, and Asia united, Australia left behind long ago: Gond-
Mugnier, J.L., Baby, P., Colletta, B., Vinour, P., Bale, P., v. 115, no. B10, https://doi.​ org/​ 10.​ 1029/​ 2010JB007443. wana Research, v. 32, p. 24–40, https://doi​.org​/10​.1016​
and Leturmy, P., 1997, Thrust geometry controlled Pickering, K.T., Carter, A., Andò, S., Garzanti, E., Limonta, /j​.gr​.2015​.02​.001.
by erosion and sedimentation; a view from analogue M., Vezzoli, G., and Milliken, K.L., 2020, Deciphering Sikder, A.M., and Alam, M.M., 2003, 2-D modelling of the
models: Geology, v. 25, no. 5, p. 427–430, https://doi​ relationships between the Nicobar and Bengal subma- anticlinal structures and structural development of the
.org​/10​.1130​/0091-7613(1997)025<0427:TGCBEA>2​ rine fans, Indian Ocean: Earth and Planetary Science eastern fold belt of the Bengal Basin, Bangladesh: Sedi-
.3.CO;2. Letters, v. 544, https://doi​.org​/10​.1016​/j​.epsl​.2020​ mentary Geology, v. 155, no. 3–4, p. 209–226, https://
Myrow, P.M., Hughes, N.C., Goodge, J.W., Fanning, C.M., .116329. doi​.org​/10​.1016​/S0037-0738(02)00181-1.
Williams, I.S., Shanchi, P., Bhargava, O.N., Parcha, Pivnik, D.A., Nahm, J., Tucker, R.S., Smith, G.O., Nyein, Simpson, G.D.H., 2010, Formation of accretionary prisms
S.K., and Pogue, K.R., 2010, Extraordinary transport K., Nyunt, M., and Maung, P.H., 1998, Polyphase influenced by sediment subduction and supplied by
and mixing of sediment across Himalayan central deformation in a fore-arc/back-arc basin, Salin sub- sediments from adjacent continents: Geology, v. 38,
Gondwana during the Cambrian–Ordovician: Geo- basin, Myanmar (Burma): AAPG Bulletin, v. 82, no. no. 2, p. 131–134, https://doi​.org​/10​.1130​/G30461​.1.
logical Society of America Bulletin, v. 122, no. 9–10, 10, p. 1837–1856, https://doi​.org​/10​.1306​/1D9BD15F- Smith, G., McNeill, L., Henstock, T.J., and Bull, J., 2012,
p. 1660–1670, https://doi​.org​/10​.1130​/B30123​.1. 172D-11D7-8645000102C1865D. The structure and fault activity of the Makran accre-
Naing, T.T., 2019, Age, depositional history, and tectonics of Polonia, A., Torelli, L., Brancolini, G., and Loreto, M.F., tionary prism: Journal of Geophysical Research: Solid
the Indo-Burman Ranges [Ph.D. thesis]: Oxford, UK, 2007, Tectonic accretion versus erosion along the Earth, v. 117, no. B07407, https://doi​.org​/10​.1029​
University of Oxford, 404 p. southern Chile trench: Oblique subduction and margin /2012JB009312.
Naing, T.T., Bussien, D.A., Winkler, W.H., Nold, M., and segmentation: Tectonics, v. 26, no. 3, https://doi​.org​/10​ Socquet, A., Vigny, C., Chamot-Rooke, N., Simons, W.,
Von Quadt, A., 2014, Provenance study on Eocene– .1029​/2006TC001983. Rangin, C., and Ambrosius, B., 2006, India and Sunda
Miocene sandstones of the Rakhine Coastal Belt, Qiu, C.G., Zhang, W.M., and Xie, X.J., 2008, Study on depo- plates motion and deformation along their boundary in
Indo-Burman Ranges of Myanmar: Geodynamic impli- sitional system in Block A4, offshore western Myan- Myanmar determined by GPS: Journal of Geophysical
cations, in Scott, R.A., Smyth, H.R., Morton, A.C., and mar: Beijing, China National Offshore Oil Corporation Research: Solid Earth, v. 111, no. B5, https://doi​.org​/10​
Richardson, N., eds., Sediment Provenance Studies in Internal Report, 170 p. .1029​/2005JB003877.

Geological Society of America Bulletin, v. 135, no. 9/10 2373

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user
Peng Zhang et al.

Stern, R.J., 2004, Subduction initiation: Spontaneous and strain across the seismogenic zone of the Kodiak shelf Yang, S.Y., and Kim, J.W., 2014, Pliocene basin-floor fan
induced: Earth and Planetary Science Letters, v. 226, and slope, Alaska: Reviews of Geophysics, v. 29, no. 3, sedimentation in the Bay of Bengal (offshore north-
no. 3–4, p. 275–292, https://doi​.org​/10​.1016​/S0012- p. 279–316, https://doi​.org​/10​.1029​/91RG00969. west Myanmar): Marine and Petroleum Geology, v. 49,
821X(04)00498-4. von Huene, R., Klaeschen, D., Cropp, B., and Miller, J., p. 45–58, https://doi​.org​/10​.1016​/j​.marpetgeo​.2013​.09​
Storti, F., and McClay, K., 1995, Influence of syntectonic 1994, Tectonic structure across the accretionary and .007.
sedimentation on thrust wedges in analogue models: erosional parts of the Japan Trench margin: Jour- Yao, W., Ding, L., Cai, F., Wang, H., Xu, Q., and Zaw, T.,
Geology, v. 23, no. 11, p. 999–1002, https://doi​.org​ nal of Geophysical Research: Solid Earth, v. 99, 2017, Origin and tectonic evolution of upper Triassic
/10 ​ . 1130 ​ / 0091-7613(1995)023<0999:IOSSOT>2​ no. B11, p. 22,349–22,361, https://doi​.org​/10​.1029​ turbidites in the Indo-Burman ranges, West Myanmar:
.3.CO;2. /94JB01198. Tectonophysics, v. 721, p. 90–105, https://doi​.org​/10​
Strasser, M., Moore, G.F., Kimura, G., Kitamura, Y., Kopf, Wang, J., Wu, F., Tan, X., and Liu, C., 2014, Magmatic evo- .1016​/j​.tecto​.2017​.09​.016.
A.J., Lallemant, S., Park, J., Screaton, E.J., Su, X., Un- lution of the Western Myanmar Arc documented by Yin, A., Dubey, C.S., Kelty, T.K., Webb, A.A.G., Harrison,
derwood, M.B., and Zhao, X., 2009, Origin and evolu- U-Pb and Hf isotopes in detrital zircon: Tectonophysics, T.M., Chou, C.Y., and Celerier, J., 2010, Geologic cor-
tion of a splay fault in the Nankai accretionary wedge: v. 612–613, p. 97–105, https://doi​.org​/10​.1016​/j​.tecto​ relation of the Himalayan Orogen and Indian Craton;
Nature Geoscience, v. 2, no. 9, p. 648–652, https://doi​ .2013​.11​.039. Part 2, Structural geology, geochronology, and tectonic
.org​/10​.1038​/ngeo609. Wang, J., Wu, F., Garzanti, E., Hu, X., Ji, W., Liu, Z., and evolution of the eastern Himalaya: Geological Soci-
Su, T., Spicer, R.A., Li, S., Xu, H., Huang, J., Sherlock, S., Liu, X., 2016, Upper Triassic turbidites of the northern ety of America Bulletin, v. 122, no. 3–4, p. 360–395,
Huang, Y., Li, S., Wang, L., Jia, L., Deng, W., Liu, J., Tethyan Himalaya (Langjiexue Group): The terminal of https://doi​.org​/10​.1130​/B26461​.1.
Deng, C., Zhang, S., Valdes, P.J., and Zhou, Z., 2019, a sediment-routing system sourced in the Gondwanide Zhang, J., Xiao, W., Windley, B.F., Wakabayashi, J., Cai,
Uplift, climate and biotic changes at the Eocene–Oligo- Orogen: Gondwana Research, v. 34, p. 84–98, https:// F., Sein, K., Wu, H., and Naing, S., 2018, Multiple
cene transition in south-eastern Tibet: National Science doi​.org​/10​.1016​/j​.gr​.2016​.03​.005. alternating forearc- and backarc-ward migration of
Review, v. 6, no. 3, p. 495–504, https://doi​.org​/10​.1093​ Wang, K., and Bilek, S.L., 2011, Do subducting seamounts magmatism in the Indo-Myanmar Orogenic Belt since
/nsr​/nwy062. generate or stop large earthquakes?: Geology, v. 39, no. the Jurassic: Documentation of the orogenic architec-
Swe, M., 2007, Geological map of Padaung–Taungup–San- 9, p. 819–822, https://doi​.org​/10​.1130​/G31856​.1. ture of eastern Neotethys in SE Asia: Earth-Science
doway region: Internal Report TF1-FL4-WM1, scale Webb, A.A.G., Yin, A., and Dubey, C.S., 2013, U-Pb zircon Reviews, v. 185, p. 704–731, https://doi​.org​/10​.1016​
1:63,000. geochronology of major lithologic units in the eastern /j​.earscirev​.2018​.07​.009.
Thein, M., and Maung, M., 2017, The Eastern (back-arc) Himalaya; implications for the origin and assembly of Zhang, J.Y., Yin, A., Liu, W.C., Wu, F.Y., Ding, L., and
Basin of central Myanmar; basement rocks, lithostrati- Himalayan rocks: Geological Society of America Bul- Grove, M., 2012, Coupled U-Pb dating and Hf isotopic
graphic units, palaeocurrents, provenance and develop- letin, v. 125, no. 3–4, p. 499–522, https://doi​.org​/10​ analysis of detrital zircon of modern river sand from
mental history, in Barber, A.J., Khin Zaw, and Crow, .1130​/B30626​.1. the Yalu River (Yarlung Tsangpo) drainage system in
M.J., eds., Myanmar: Geology, Resources and Tecton- Webb, M., Gough, A., Vannucchi, P., Lünsdorf, N.K., and southern Tibet; constraints on the transport processes
ics: Geological Society, London, Memoir 48, p. 169– McNeil, J., 2021, Sedimentary provenance of the Plio- and evolution of Himalayan rivers: Geological Society
183, https://doi​.org​/10​.1144​/M48​.8. Pleistocene Nicobar Fan: Complex sourcing revealed of America Bulletin, v. 124, no. 9–10, p. 1449–1473,
Tobin, H.J., and Saffer, D.M., 2009, Elevated fluid pressure through Raman spectroscopy heavy mineral analysis: https://doi​.org​/10​.1130​/B30592​.1.
and extreme mechanical weakness of a plate bound- Marine and Petroleum Geology, v. 125, https://doi​.org​ Zhang, P., Mei, L., Hu, X., Li, R., Wu, L., Zhou, Z., and
ary thrust, Nankai Trough Subduction Zone: Geol- /10​.1016​/j​.marpetgeo​.2020​.104874. Qiu, H., 2017, Structures, uplift, and magmatism
ogy, v. 37, no. 8, p. 679–682, https://doi​.org​/10​.1130​ Westerweel, J., Roperch, P., Licht, A., Dupont-Nivet, G., of the Western Myanmar Arc: Constraints to mid-
/G25752A​.1. Win, Z., Poblete, F., Ruffet, G., Swe, H.H., Thi, M.K., Cretaceous–Paleogene tectonic evolution of the
Tun, N., Kyaw, M.M., Win, K.M., and Hlaing, K., 2007, and Aung, D.W., 2019, Burma Terrane part of the western Myanmar continental margin: Gondwana
Prospect summary of offshore block A7: Ministry of Trans-Tethyan arc during collision with India accord- Research, v. 52, p. 18–38, https://doi​.org​/10​.1016​/j​
Energy, Myanma Oil and Gas Enterprise Internal Re- ing to palaeomagnetic data: Nature Geoscience, v. 12, .gr​.2017​.09​.002.
port, 90 p. no. 10, p. 863–868, https://doi​.org​/10​.1038​/s41561- Zhang, P., Najman, Y., Mei, L., Millar, I., Sobel, E.R., Carter,
United Nations, 1979, Geology and exploration geochem- 019-0443-2. A., Barfod, D., Dhuime, B., Garzanti, E., Govin, G.,
istry of part of the Northern and Southern Chin Hills Westerweel, J., Licht, A., Cogné, N., Roperch, P., Dupont Vezzoli, G., and Hu, X., 2019b, Palaeodrainage evo-
and Arakan Yoma, Western Burma: New York, United Nivet, G., Kay Thi, M., Swe, H.H., Huang, H., Win, lution of the large rivers of East Asia, and Himala-
Nations Development Programme Technical Report 4, Z., and Wa Aung, D., 2020, Burma Terrane collision yan-Tibet tectonics: Earth-Science Reviews, v. 192,
DP/UN/BUR-72-002/13, 59 p. and northward indentation in the Eastern Himalayas p. 601–630, https://doi​.org​/10​.1016​/j​.earscirev​.2019​
Vadlamani, R., Wu, F., and Ji, W., 2015, Detrital zircon recorded in the Eocene–Miocene Chindwin Basin .02​.003.
U–Pb age and Hf isotopic composition from foreland (Myanmar): Tectonics, v. 39, no. 10, https://doi​.org​/10​ Zhang, P., Mei, L., Jiang, S., Xu, S., Donelick, R.A., Li, R.,
sediments of the Assam Basin, NE India: Constraints .1029​/2020TC006413. and Zhang, H., 2021a, Erosion and sedimentation in
on sediment provenance and tectonics of the Eastern Xiong, Z., Ding, L., Spicer, R.A., Farnsworth, A., Wang, X., SE Tibet and Myanmar during the evolution of the Bur-
Himalaya: Journal of Asian Earth Sciences, v. 111, Valdes, P.J., Su, T., Zhang, Q., Zhang, L., Cai, F., Wang, mese continental margin from the Late Cretaceous to
p. 254–267, https://doi​.org​/10​.1016​/j​.jseaes​.2015​.07​ H., Li, Z., Song, P., Guo, X., and Yue, Y., 2020, The ear- Early Neogene: Gondwana Research, v. 95, p. 149–175,
.011. ly Eocene rise of the Gonjo Basin, SE Tibet: From low https://doi​.org​/10​.1016​/j​.gr​.2021​.04​.005.
Vermeesch, P., 2004, How many grains are needed for a prov- desert to high forest: Earth and Planetary Science Let- Zhang, X., Chung, S., Lai, Y., Ghani, A.A., Murtadha, S.,
enance study?: Earth and Planetary Science Letters, ters, v. 543, https://doi.​ org/​ 10.​ 1016/​ j.​ epsl.​ 2020.​ 116312. Lee, H., and Hsu, C., 2019a, A 6000-km-long Neo-
v. 224, no. 3–4, p. 441–451, https://doi​.org​/10​.1016​/j​ Xu, Z., Wang, Q., Cai, Z., Dong, H., Li, H., Chen, X., Duan, Tethyan arc system with coherent magmatic flare-ups
.epsl​.2004​.05​.037. X., Cao, H., Li, J., and Burg, J., 2015, Kinematics of and lulls in South Asia: Geology, v. 47, no. 6, p. 573–
Vermeesch, P., 2013, Multi-sample comparison of detrital the Tengchong Terrane in SE Tibet from the late Eocene 576, https://doi​.org​/10​.1130​/G46172​.1.
age distributions: Chemical Geology, v. 341, p. 140– to early Miocene: Insights from coeval mid-crustal de- Zhang, X., Chung, S., Tang, J., Maulana, A., Mawaleda, M.,
146, https://doi​.org​/10​.1016​/j​.chemgeo​.2013​.01​.010. tachments and strike-slip shear zones: Tectonophysics, Oo, T., Tien, C., and Lee, H., 2021b, Tracing Argo-
Vermeesch, P., 2018, IsoplotR: A free and open toolbox for v. 665, p. 127–148, https://doi​.org​/10​.1016​/j​.tecto​.2015​ land in eastern Tethys and implications for India-Asia
geochronology: Geoscience Frontiers, v. 9, p. 1479– .09​.033. convergence: Geological Society of America Bulletin,
1493, https://doi​.org​/10​.1016​/j​.gsf​.2018​.04​.001. Yang, L., Xiao, W., Jalalur Rahman, M.J., Windley, B.F., v. 133, no. 7–8, p. 1712–1722, https://doi​.org​/10​.1130​
Vigny, C., Socquet, A., Rangin, C., Chamot-Rooke, N., Schulmann, K., Ao, S., Zhang, J.E., Chen, Z., Hos- /B35772​.1.
Publlier, M., Bouin, M.N., Bertrand, G., and Becker, sain, M.S., and Dong, Y., 2020, Indo-Burma passive Zhao, R.M., Zhao, H.X., Zhu, G.H., Yang, S.L., Jiang, Y.,
M., 2003, Present-day crustal deformation around amalgamation along the Kaladan Fault: Insights from Xie, N., Lv, M., and Xie, X.J., 2006, General evalua-
Sagaing Fault, Myanmar: Journal of Geophysical Re- zircon provenance in the Chittagong-Tripura Fold Belt tion and well location proposal for Block M, onshore
search: Solid Earth, v. 108, no. B11, https://doi​.org​/10​ (Bangladesh): Geological Society of America Bulletin, western Myanmar: Beijing, China National Offshore
.1029​/2002JB001999. v. 132, no. 9–10, p. 1953–1968, https://doi​.org​/10​.1130​ Oil Corporation Internal Report, 144 p.
von Huene, R., and Klaeschen, D., 1999, Opposing gradi- /B35429​.1. Zheng, Q.G., 2008, Sedimentary facies study and reservoir
ents of permanent strain in the aseismic zone and elastic Yang, S.L., Zhao, R.M., Zhao, H.X., Zhu, G.H., Xie, N., Ji- prediction in a compressional tectonic area, Block A4,
strain across the seismogenic zone of the Kodiak shelf ang, Y., and Cai, W.J., 2007, Research achievements of offshore western Myanmar: Beijing, China University
and slope, Alaska: Tectonics, v. 18, no. 2, p. 248–262, Block M, onshore western Myanmar: Beijing, China of Geosciences Internal Report, 62 p.
https://doi​.org​/10​.1029​/1998TC900022. National Offshore Oil Corporation Internal Report,
von Huene, R., and Ranero, C.R., 2003, Subduction erosion 167 p. Science Editor: Wenjiao Xiao
and basal friction along the sediment-starved conver- Yang, S.L., Zhu, G.H., Zhao, R.M., Zhao, H.X., Jiang, Y., Associate Editor: Kathryn Cutts
gent margin off Antofagasta, Chile: Journal of Geo- Xie, N., and Cai, W.J., 2009, Analysis of petroleum
Manuscript Received 25 April 2022
physical Research: Solid Earth, v. 108, no. B2, https:// geological conditions and well location target selec- Revised Manuscript Received 9 September 2022
doi​.org​/10​.1029​/2001JB001569. tion in Block A4, offshore western Myanmar: Beijing, Manuscript Accepted 9 October 2022
von Huene, R., and Scholl, D.W., 1991, Opposing gradients China National Offshore Oil Corporation Internal Re-
of permanent strain in the aseismic zone and elastic port, 295 p. Printed in the USA

2374 Geological Society of America Bulletin, v. 135, no. 9/10

Downloaded from http://pubs.geoscienceworld.org/gsa/gsabulletin/article-pdf/135/9-10/2348/5944484/b36560.1.pdf


by University of Tasmania user

You might also like